<<

This manuscript is contextually identical with the following published paper:

Kuszegi, G., Siller, I., Dima, B., Takács, K., Merényi, Zs., Varga, T., Turcsányi, G., Bidló, A.,

Ódor, P. 2015. Drivers of macrofungal species composition in temperate forests, West

Hungary: functional groups compared. Fungal Ecology 17: 69-83. DOI:

10.1016/j.funeco.2015.05.009

The original published pdf available in this website:

http://authors.elsevier.com/sd/article/S0378112713004295

Title: Drivers of macrofungal species composition in temperate forests, West Hungary: functional groups compared

Authors: Gergely Kutszegi1,*, Irén Siller2, Bálint Dima3, 6, Katalin Takács3, Zsolt Merényi4,

Torda Varga4, Gábor Turcsányi3, András Bidló5, Péter Ódor1

1MTA Centre for Ecological Research, Institute of Ecology and , Alkotmány út 2–4,

H-2163 Vácrátót, Hungary, [email protected], [email protected].

2Department of Botany, Institute of Biology, Szent István University, P.O. Box 2, H-1400

Budapest, Hungary, [email protected].

3Department of Nature Conservation and Landscape Ecology, Institute of Environmental and

Landscape Management, Szent István University, Páter Károly út 1, H-2100 Gödöllő,

Hungary, [email protected], [email protected], [email protected].

4Department of Plant Physiology and Molecular Plant Biology, Eötvös Loránd University,

1

Pázmány Péter sétány 1/C, H-1117 Budapest, Hungary, [email protected], [email protected].

5Department of Forest Site Diagnosis and Classification, University of West-Hungary, Ady út

5, H-9400 Sopron, Hungary, [email protected].

6Department of Biosciences, University of Helsinki, P.O. Box 65, 00014 University of

Helsinki, Helsinki, Finland, [email protected].

* corresponding author, Fax: +36 28 360110, email address: [email protected] (Gergely Kutszegi)

Footnote to the title: Environmental drivers of macrofungal species composition

2

Abstract

The most influential environmental drivers of macrofungal species composition were studied in managed, even-aged, mixed forests of Őrség National Park, Hungary. Functional groups of macrofungi were analyzed separately by non-metric multidimensional scaling and redundancy analysis exploring their relations to tree species composition, stand structure, soil/litter conditions, microclimate, landscape, and management history. Some evidence was given that macrofungi are related to drivers that are relatively easy to measure. It was found that wood- inhabiting fungal species composition is driven primarily by the species composition of living trees, while substrate properties and microclimate play minor roles. The terricolous saprotrophic community was determined principally by a litter pH gradient involving tree species composition and soil/litter properties. Microclimate had no concordant effect. No obvious underlying gradients were detected on ectomycorrhizal fungal species composition; however, tree size and litter pH had significant effects. For each group, no clear responses to landscape or management history were detected.

Key words: biodiversity; ectomycorrhizal fungi; environmental variation; fungal community gradients; host specificity; soil properties; sampling; terricolous saprotrophic fungi; wood-inhabiting fungi

3

Introduction

Forest-dwelling macrofungal assemblages have been classified into three main functional groups: wood-inhabiting (including wood saprotrophs and necrotrophic parasites), ectomycorrhizal (EcM) and terricolous saprotrophic communities (Winterhoff, 1992). In a global perspective, an enormous volume of research has been reported on the responses of macrofungal community composition to environmental variation. It was revealed that wood- inhabiting fungi are driven principally by the amount and diameter (Heilmann-Clausen and

Christensen, 2004; Sippola et al., 2005; Ódor et al., 2006; Lonsdale et al., 2008), decay stage

(Heilmann-Clausen and Christensen, 2003b; Siller, 2004; Heilmann-Clausen et al., 2014), age

(Heilmann-Clausen, 2001), species identity (Sippola et al., 2005; Küffer et al., 2008), complexity (Heilmann-Clausen and Christensen, 2003a), and spatio-temporal availability

(Siitonen, 2001; Bässler et al., 2010; Halme et al., 2013) of dead wood. The microclimatic variation and pH within the wood (Boddy, 1992, 2001; Salerni et al., 2002) or the interactions with other organisms (van der Wal et al., 2013) had also significant effects. EcM community compositions were found to be structured strongly by the N content (Toljander et al., 2006;

Cox et al., 2010; Suz et al., 2014), pH (Baar and ter Braak, 1996; Talbot et al., 2013) as well as temperature and moisture of soil (Claridge et al., 2000; Jones et al., 2003), species composition of host trees (Kernaghan et al., 2003; Smith and Read, 2008; Morris et al., 2009), season (over the course of even a month) (Courty et al., 2008), fungal dispersal limitation among host trees (Peay et al., 2010), and timing of colonization and interspecific competition on the root surface (Kennedy et al., 2009; Kennedy, 2010). In the same context, little is known about the determinants of terricolous saprotrophic communities, but the effects of litter quantity and pH (Tyler, 1991; Ferris et al., 2000; Talbot et al., 2013), P content of the soil

(Reverchon et al., 2010), tree species composition (O’Hanlon and Harrington, 2012), and temperature (McMullan-Fisher et al., 2009) were documented to be highly important.

4

Many influential environmental drivers have been revealed, but are there drivers with consistent importance on macrofungal functional groups? When such drivers are sought, many difficulties are encountered. The relative importance of drivers varies across spatial scales (Claridge et al., 2000; Lilleskov and Parrent, 2007; Büntgen et al., 2012) and along environmental gradients, such as elevation (Gómez-Hernández et al., 2012; Sundqvist et al.,

2013) and rainfall (Lindblad, 2001; Salerni et al., 2002). Also, the relative effects of drivers can be biased strongly by the edaphic heterogeneity of the studied habitats, and the actually limiting factors (resources or environmental conditions) in a habitat can have a disproportionately high influence on species composition (McMullan-Fisher, 2008). In addition, community level responses are difficult to reveal, since great species diversity is found within fungal communities in which each species have slightly different environmental requirements (Boddy et al., 2008).

Based on the studies mentioned in the first paragraph, our knowledge of fungal community responses to environmental variation is biased by research history: (1) the majority of studies have been conducted in Northern or Western or in North America, thus, large regions are still underrepresented; (2) the studies have rarely been focused on more than two functional groups (except e.g. Humphrey et al., 2000; Sato et al., 2012); (3) to obtain a clearer picture, many authors have used a limited pool of environmental factors and hence, several environmental impacts with probable significant effects remained unexplored on the sampling sites.

Given this complexity and research gaps, the present study has been designed in even- aged, managed forests with a restricted number of habitat types to try to reduce the effects of edaphic heterogeneity. By including several variables suggested by the literature, other factors that characterize the landscape and management history were also examined.

In accordance with the studies referenced in the first paragraph, it can be hypothesized

5

that (1) the effects of substrate properties, tree species composition, and microclimate have the strongest effects on macrofungal species composition at a stand scale, and (2) the relative influence of these factors differs among wood-inhabiting, EcM, and terricolous saprotrophic communities. The aims of this study are to find the most important environmental factors that best explain the macrofungal species composition of wood-inhabiting, EcM and terricolous saprotrophic communities, and provide information on the environmental requirements of fungal species.

6

Materials and methods

Study area

This study has been carried out in Őrség National Park (ŐNP), West Hungary (46°

51’–55’ North, 16° 06’–24’ East (Fig 1a). In the ŐNP, the precipitation ranges between 700 and 800 mm yearly. Between 1901 and 2000, the mean minimum and maximum temperatures in winter were respectively –7.4 and 6.0°C, while in summer 13.5 and 23.8°C (measured in a nearby town, Szombathely, Hungarian Meteorological Service, OMSZ). The landscape is divided into hills and wide valleys at the elevation range of 250–350 m above sea level. The bedrock consists of alluvial gravel and clay. Nutrient-poor brown forest soils with pseudogley or lessivage (planosols or luvisols) are the most frequent soil types (Halász, 2006; Dövényi,

2010). The pH of the soil is acidic; it tends to range from 4.0 to 4.8 with a mean of 4.3 (Juhász et al., 2011).

Presently, forests cover 80% of the ŐNP region, which has an area of ca. 350 km2

(Dövényi, 2010). Stands are dominated by ( L.), sessile and pedunculate oak [Quercus petraea (Matuschka) Liebl. and Q. robur L.], hornbeam (Carpinus betulus L.), and Scots pine (Pinus sylvestris L.). Forests are sometimes monodominant, but more often form mixed stands with great compositional diversity. The most frequent non-dominant tree species are Betula pendula Roth, Picea abies (L.) Karst., Populus tremula L., Castanea sativa

Mill., Prunus avium L., Tilia spp., and Acer spp. (Tímár et al., 2002). ŐNP is characterized by the highest proportion of private forest stands in Hungary where the dominant tree species usually varies from stand to stand. Therefore, the ŐNP is a suitable region for studying the effects of tree species on macrofungal communities.

Between the 12th and 19th centuries, the landscape was characterized by a rotation cycle in land use: small pieces of forests, meadows and arable lands were replaced by each other. Meanwhile all the pristine forests were cut. Leaf-litter was collected widely in the

7

secondary stands and used as bedding for farm animals. A specific ridge planting system was applied on the arable lands to decrease the high levels of groundwater in the upper soil layers; plants were set onto the top of the ridges. The nutrient-poor arable lands had to be fallowed often for many years whilst they were frequently regrown by pine and ; slash and burn was used to return the regenerated forests to arable land uses. As a consequence of these activities, the region was characterized by much soil erosion, leaching and acidification. Due to that, the proportion of pioneer trees (Pinus sylvestris and Betula pendula), acidofrequent herbs, bryophytes and lichens increased. Now, these traditional cultivation practices have ceased. Currently, a spontaneous stem selection method in the private forests and a shelterwood management with a rotation period of 70–110 years in the state forests are applied. As a result of this, an increasing proportion of deciduous trees and mesophytic herbs can be observed in the region (Gyöngyössy, 2008).

Environmental data collection

Similar habitats without strong effects of edaphic heterogeneity (that would make the environmental data noisy) are required for finding the environmental factors of forest macrofungi that drive their species composition. In accordance, forest stands were selected by a stratified random sampling based on the Hungarian Forestry Database (Hungarian Central

Agricultural Office, Forestry Directorate, www.nebih.gov). The even-aged, 70–100 years old, spatially independent stands (the minimum distance is 500 m between them) chosen are located in relatively flat areas and not influenced directly by surface waters. These stands were grouped based on the most frequent tree species. Thirty-five stands were selected randomly from these groups representing a gradient along the characteristic tree species combinations of the region. A 40 m × 40 m plot was assigned in each selected stand.

Geographical positions of plots are shown in Fig 1b; GPS coordinates are available in Siller et

8

al. (2013). The plots are scattered in a 160 km2 area. In the middle of each plot, a 30 m × 30 m sampling unit was assigned for macrofungi surveys. Sampling units were divided into thirty- six 5 m × 5 m quadrats arranged systematically.

Environmental data that are easy to measure on the sites were used as potential explanatory variables to explain the species composition of macrofungal communities. Fifty- two variables representing tree species composition, stand structure, soil and litter conditions, microclimate, landscape structure, and management history were measured (Table 1).

Tree species composition was expressed based on the relative volume of tree species by merging all taxa within the same , e.g. oaks (Quercus cerris, Q. petraea, Q. robur) and limes (Tilia cordata, T. platyphyllos). Volume of tree individuals was computed by species specific equations using the height and diameter of trees at breast height (DBH) (Sopp and Kolozs, 2000). Shannon diversity of tree species was calculated based on relative tree volumes and using natural logarithm (Shannon and Weaver, 1949).

Regarding stand structure, each tree within the 40 m × 40 m plots and larger than 5 cm

DBH was mapped; tree species identity, DBH and height were recorded. Coarse woody debris

(CWD) longer than 50 cm and thicker than 10 cm, and snags (including stumps) thicker than

5 cm were measured and mapped; volumes were computed by assuming that they were cylinders. Decay stage of CWD was determined according to Ódor and van Hees (2004).

Projected onto the soil surface, the relative area covered by woody debris [fine (FWD) and coarse units together], litter, bare soil, bryophytes, and understory vegetation (including herbs and seedlings shorter than 50 cm) were estimated visually in the 5 m × 5 m quadrats; and their results were transformed into m2/ha. Shrub density was measured by counting each arboreal individual (including regenerating trees) thinner than 5 cm DBH and taller than 50 cm.

Soil and litter conditions were measured within the sampling units by sampling five points arranged systematically. Litter was collected from 30 cm × 30 cm areas. Soil cubes of

9

15 cm × 15 cm were sampled from the vertical layer of 0–10 cm. Soil and litter pH were measured potentiometrically by a pH meter in the supernatant suspension of the sample.

Determination of hydrolytic (y1) and exchangeable (y2) acidity were carried out by titration

3 3 with NaOH; soil samples were extracted by 1 mol/dm Ca(CH3COO)2 and 1 mol/dm KCl solutions, respectively (Bellér, 1997). The organic C and total N content of soil and litter were measured according to ISO (1995, 1998) applying dry combustion elementary analysis by

Elementar vario EL III CNS equipment. The P and K contents of the soil were extracted by an ammonium lactate/acetic acid solution based on Bellér (1997).

Air humidity and temperature measurements were conducted in the center of each sampling unit at 1.3 m height using Voltcraft DL-120 TH data loggers. For both measurements, dissimilarity values were calculated between the measured values of two nearby reference sites and the measured values of the studied sampling units. Measurements were synchronized in time and lasted for 24 hours by setting five minutes recording frequency. By repeating the same procedure, eight measurements were carried out in different months of the vegetation periods between 2009 and 2011, and their results were averaged.

Relative diffuse light was measured by LAI-2000 Plant Canopy Analyzer in the center of each sampling unit at 1.3 m height and always at dusk (Tinya et al., 2009).

The proportion of landcover types (forests, permanently open patches and cutting areas) was calculated inside a circle of 300 m radius surrounding each plot. Measurements were carried out using aerial photographs and topographic maps. Stands older than 20 years were considered to be forests; younger ones were defined as cutting areas. Landscape diversity was expressed by the Shannon diversity index based on the relative cover of landscape elements (Shannon and Weaver, 1949).

Management history was demonstrated based on the map made by the Habsburg

Empire in 1853 during the Second Military Survey (Arcanum, 2006). According to this map,

10

the same landscape variables were computed that were used for characterizing the recent landscape. Historical land use types of sampling units were fixed as binary variables.

Fungi data

By reason of the large total area (31,500 m2) of sampling units, sporocarp surveys were conducted to characterize the macrofungal species composition. Macrofungal surveys sampled basidiomycetes (excluding most of the resupinate non-poroid taxa) and ascomycetes that develop sporocarps visible to the naked eye (larger than 2 mm). Sporocarps were sampled three times in each sampling unit: in August 2009, May 2010 and during September–

November 2010. The precipitation in 2010 was far above average resulting in high sporocarp production in the region. Thus, the duration of the third survey was relatively long: 48 days between 19 of September and 5 of November (early November is generally the very end of the main fruiting period in Hungary). Dried specimens were deposited in the Hungarian

Natural History Museum, Department of Botany (BP), Budapest.

To obtain presence-absence data for macrofungi, the species identity of taxa was recorded in each quadrat of each sampling unit in each sampling period. Accordingly, the total of the number of times a species was found in a quadrat in a sampling unit was a calculated abundance measure (a local frequency value) for each collected . The maximum value of the local frequency of a species is 36 × 3 = 108 based on the 36 quadrats in a sampling unit and the three sampling periods. Therefore, our community data form a multivariate matrix of fungal species and sampling units where species performance was expressed by local frequency values. The total number of sampling units occupied by a fungus was also calculated (Supplementary Table 1).

Species identification procedures are detailed in Siller et al. (2013). The identity and nomenclature of sampled taxa were determined by using monographs, books and papers.

11

MycoBank (www.mycobank.org, assessed between 19 and 20 of April 2013) and more rarely

Knudsen and Vesterholt (2012) were used to verify up-to-date scientific names and author(s) of fungal species.

The macrofungal taxa were classified into three main functional groups: terricolous saprotrophic fungi living on litter or any kind of buried plant debris in the uppermost 10 cm of the soil; wood-inhabiting fungi colonizing dead branches, twigs, logs or snags on the ground, and trunks or roots of living wood; and EcM fungi representing a well definable, stand-alone group (Tedersoo et al., 2010).

Data analyses

Environmental variables that drove the species composition of wood-inhabiting, terricolous saprotrophic and EcM fungi were examined separately for each functional group by two different ordination methods: redundancy analysis (RDA) and non-metric multidimensional scaling (NMDS). Results of both methods are evaluated in this study by looking for consistency in their environmental interpretations, but more focus was put on the

NMDS results because these models had better explanatory powers.

RDA plots points of species and sampling units in a space defined by the environmental variables, and was used here to represent the best fit of species abundances to the environmental data. This method is a constrained ordination based on a model of linear species response to the underlying environmental gradient where the rare species (listed in

Supplementary Table 1 and found on less than four sampling units) are often dropped from the analysis, as it was also done in this work to satisfy its requirements (Legendre and

Legendre 1998). RDA was chosen as a suitable direct gradient analysis after the detection of short gradient lengths (2–3 SD units) revealed by the detrended correspondence analyses of functional groups (Lepš and Šmilauer, 2003). RDA models were built by the manual forward

12

selection of explanatory variables and testing the effects of variables on community data by F- statistics applying Monte Carlo simulations with 999 permutations; significance of all canonical axes were tested similarly (ter Braak and Šmilauer, 2002). Log-transformed local frequency values of taxa were used for RDAs.

By contrast, NMDS is an unconstrained ordination that avoids the assumption of linear relationships among variables and provides a valuable representation of the overall community structures without constricting the analysis to the frequent species only. In this regard, NMDS is a powerful tool, but it is not designed principally for finding the most important environmental drivers of species composition. That is, the environmental interpretation of the NMDS results can be achieved by fitting vectors subsequently onto the

NMDS solutions, which were reached independently from the environmental data (Oksanen,

2013). In this study, NMDS was carried out following McCune & Grace (2002) and Oksanen

(2013). Regarding each functional group, “local” NMDS model (Sibson, 1972) was fitted where an independent monotonic regression is used for each sampling unit in contrast to the

“global” NMDS model (Kruskal, 1964), which is fitted from a global point of view on ranked dissimilarities. According to Prentice (1977), local NMDS can be more suitable for evaluating ecological gradients than the global NMDS model because it is sensitive to the local environment of each point in the ordination space supposing that the environment itself can change along a gradient. NMDS was run on Sørensen (Bray–Curtis) distances and it obtained a much stronger description of community structures compared to the other tested distance methods: “Jaccard”, “Canberra” and “Euclidean”. Random starting configurations (20 for each functional group) were used for finding the best stable solutions. The dimensionality of each studied community data was revealed based on the Supplementary Fig 1–3e. The

Kendall’s rank correlation coefficients (τ) were calculated between the original distance matrices and the ordination distances, and they were plotted against the final stress values

13

testing the dimensions between 1 and 10. Three dimensional solutions were chosen to be plotted in this study. NMDS stress was measured by Kruskal’s stress formula 1 multiplied by

100 (Kruskal, 1964). For representing goodness of fit, Shepard diagrams and the best-fit monotonic regressions of distances were plotted in Supplementary Fig 1–3c and d. The environmental variables fitted significantly (p < 0.05) onto the NMDS solutions were screened for strong (|r| > 0.5) collinearities and intercorrelated ones with a weaker relationship to the response variables were removed.

Before the analyses, a preparative procedure was completed for the environmental variables: (1) their normality was checked and, if needed, ln-transformation was applied

(Table 1), and (2) they were centred and standardized by standard deviation. It was supposed that our community data is biased by the third, 48-day long sporocarp survey during which the vast majority of records were obtained and the field visit of some sampling units were extended to the very end (or longer) of the fruiting period of some species. Therefore, the days of this sampling period were numbered from 1 to 48 and a “sampling time” variable was created. Sampling time correlated often strongly with any of the ordination axes regarding each functional group (ranges of |r| and p-values: 0.624–0.800; 0.003–0.001). Geographical longitude and/or latitude coordinates of sampling units also have strong correlations with the response variables (|r| = 0.534–0.642; p = 0.013–0.002). Moreover, unexpectedly, these three variables (sampling time, latitude and longitude coordinates) and some of the studied environmental variables were also related: |r| = 0.402–0.493; p =0.014–0.003. Thus, the amount of variation that can be attributed exclusively to the effects of sampling time and geographical coordinates was measured by applying partial regression analysis according to

Legendre and Legendre (1998); the residuals of the partial regression models were used for further analyses. These corrected environmental variables were fitted onto the NMDS solutions, while the RDA models (with the ability to use corrected variables) were built by

14

using the original environmental variables and entering the geographical coordinates and sampling time as covariates on each occasion.

R for Windows 3.0.1 (R Core Team, 2013) and, if required, the R package “vegan” v.2.0-8 (Oksanen et al., 2013) was used for carrying out preliminary tests of environmental variables, correlations, partial regressions, and NMDS. The R package “Rcmdr” v.2.1-4 (Fox,

2005) was applied for displaying spinning 3-D NMDS solutions. Canoco for Windows 4.5

(ter Braak and Šmilauer, 2002) was applied for RDAs.

15

Results

Fungal diversity

In this study, 687 macrofungal taxa were collected and identified (Supplementary

Table 1). Taxa belonging to phylum (631 species, 167 genera) were more species rich over ascomycetous taxa (56 species, 29 genera). A total of 13,396 records and

1,556 specimens were obtained. The vast majority of records (11,647 pieces, 87%) were collected during the third field survey in autumn 2010, whereas the total number of records was 1,313 (10%) in August 2009 and 436 (3%) in May 2010. Macrofungal taxa were classified into eight functional groups (Supplementary Table 1). The three most species rich functional groups were studied, in which a few abundant and a large number of rare species were found (Table 2).

Tree species identity drove wood-inhabiting fungi

Thirty-two taxa (out of 245) were found in 14 or more sampling units. These taxa are plotted in Fig 2a (displaying the NMDS results) and in Supplementary Fig 4 (showing the

RDA plot). The explanatory powers and statistical reliability of the six variables fitted onto the final NMDS solution are detailed in Table 3. In brief, concordant results were revealed by each NMDS run (including all tested distance methods and dimensionality) and RDA: tree species composition (the relative volumes of dominant tree species) had the strongest effect on wood-inhabiting community composition. In Fig 2a, axis 1 represented 17.2% of the variation and was correlated highly with the relative volume of oaks and the hydrolytic acidity of the soil. Axis 2 (9.4% variation) showed a strong correlation with the relative volume of beech, while axis 3 (6.9% variation) was related to the species richness of trees, the relative volume of , and the total cover of dead wood. Unexpectedly, dead wood properties had no significant effects in RDA. Both methods pointed out that (1) the high proportions of

16

deciduous (mainly beech and oak) trees on the sampling units were preferred by the majority of fungal species, (2) the relative volumes of tree species defined a clear deciduous– coniferous gradient in the ordination diagrams, and (3) there was no significant effect of the surrounding landscape on the species composition of wood-inhabiting fungi. The effects of air temperature and the historical proportion of meadows were significant based on the RDA only.

Both methods revealed very similar environmental requirements for the most frequent fungal taxa. The following fungal species were strongly associated with beech stands:

Antrodiella fragrans, nummularia, Hypoxylon fragiforme, Mycetinis alliaceus,

Polyporus varius, Postia subcaesia, Skeletocutis nivea, Trametes versicolor, carpophila and X. hypoxylon. The species B. nummularia, T. versicolor, and X. hypoxylon were also correlated with more neutral litter pH and trees with larger mean DBH. Wood- inhabiting fungi in oak-dominated stands with higher air temperature and higher soil hydrolytic acidity were Hymenochaete rubiginosa, Schizopora paradoxa s.l., Stereum ochraceoflavum and S. subtomentosum. Common species in coniferous (mainly pine- dominated) stands with higher total cover of dead wood and lower air temperature were

Mycena epipterygia and Ramaria stricta (the relative volume of conifers in the NMDS plot and the relative volume of Scots pine in the RDA were highly correlated: |r| = 0.964, p <

0.001) indicating that the two variables have a very similar effect on wood-inhabiting fungi in the sampling units.

A pH gradient structured terricolous saprotrophic fungi

One hundred and twenty-seven taxa were found out of which 12 species occupied more than 14 sampling units. Fig 2b shows the optimal positions of these species in the final

NMDS solution, while Supplementary Fig 5a demonstrates the RDA plot. All of the 35

17

sampling units were examined by RDA, whereas NMDS was run omitting the four sampling units with zero or very low counts of terricolous saprotrophic fungi. Four variables were fitted significantly onto the NMDS solution; their explanatory powers and statistical reliability are available in Table 3. Broadly speaking, both methods gave similar results. Terricolous saprotrophic community composition was driven principally by a definite litter pH gradient along two environmental variables: Scots pine proportion and the pH of litter. The same results were obtained from all NMDS runs where the other tested distance methods and different dimensionality were applied. Here, NMDS axis 1 represented 35.4% of the variation and was not correlated strongly with any of the environmental variables. Axis 2 (9.1% variation) was correlated highly with the relative volume of Scots pine, the pH of litter, and the density of large trees. The K content of the soil had high scores along axes 2 and 3. By contrast, RDA highlighted two other variables: the mean daily air temperature and the N content of the soil as being of great importance. In general, both of these factors were correlated negatively with the whole fungal community (Supplementary Fig 5b). No significant relations were detected by either RDA or NMDS with respect to the historical forest management or the surrounding landscape.

Concerning the environmental requirements of the frequent species, both methods supported Auriscalpium vulgare, Baeospora myosura, and molle to be common elements of pine-dominated stands with a low litter pH and a low density of large trees. The positions of other frequent taxa in the two ordination diagrams were rather unstable, but

Mycena sanguinolenta and butyracea were always found to be unrelated to the pH gradient.

EcM fungi: no obvious gradients detected

Altogether 290 EcM taxa were identified. Thirty of them were frequent, and collected

18

in more than 14 sampling units. These species are displayed in the NMDS (Fig 2c) and the

RDA (Supplementary Fig 6) diagrams. Four variables were fitted significantly onto the final

NMDS solution (for details of fit see Table 3). RDA revealed EcM fungi as a mainly host restricted functional group with the strongest effects being beech proportion and the mean

DBH of trees, while NMDS detected the density of large trees (with the highest influence) and substrate related factors (the relative volume of decayed logs, litter pH, and soil P content) to be important drivers of EcM fungal species composition. Litter pH and tree size

(mean DBH in RDA and large trees in NMDS) were important by both methods. When

NMDS was run with the other tested dimensionality and distance methods, it returned concordant results. In Fig 2c, NMDS axis 1 explained 33.8% of the variation and was not related strongly to any of the environmental variables. Axis 2 (15.6% variation) correlated highly with the P content of the soil, while axis 3 (4.4% variation) was related strongly to the relative volume of decayed logs, the density of large trees, and the pH of litter. Using RDA, three less important environmental factors were also significant: the proportion of forests in the landscape, the mean relative diffuse light, and the Shannon diversity of landscape elements. No obvious underlying gradients (supported by more than one fitted variable) were detected by either method.

Regarding fungal species, the relative volume of beech in the NMDS model, however, had no significant effect on the whole EcM community, but the optima positions of beech- dominated sampling units were close to the species Lactarius blennius, Pseudocraterellus undulatus, Russula emetica, and Tricholoma ustale (data not shown). Except P. undulatus, these species also were associated with high relative volumes of decayed logs. RDA, more or less, underlined these results revealing three more species (Inocybe petiginosa, Lactarius subdulcis, and Tricholoma sulphureum) as beech associated ones. Here, I. petiginosa and P. undulatus preferred closed canopy conditions. For both methods, the placements of Amanita

19

rubescens, Clavulina coralloides, Lactarius quietus, Russula heterophylla, R. nigricans, and

R. undulata in the ordination space were similar (they were close to each other), but RDA revealed them as characteristic taxa of open stands with more light and lower pH of litter, while NMDS emphasized strong relationships between these species and the high P content of the soil. In both models, mainly oak-dominated stands were situated close to these taxa.

20

Discussion

Fungal diversity and drivers of frequent taxa

In this work, 687 macrofungal species were recorded in total during only three sporocarp surveys and by studying a restricted number of habitat types. Altogether 30 taxa were obtained with clear concordant responses to the environment based on both the NMDS and RDA diagrams. In general, these results agreed with the findings of other studies discussed in Table 4.

Tree species composition

It was shown that the species composition of trees has the highest relevance on wood- inhabiting fungal species composition at a scale of forest stands (Fig 2a). Comparative studies in Europe have also identified tree species composition as having determinative effects on wood-inhabiting fungal species composition (e.g. Humphrey et al., 2000; Sippola et al., 2005;

O’Hanlon and Harrington, 2012). In the present study, a clear distinction between coniferous and deciduous tree species was found, which has been confirmed also by other studies (Küffer et al., 2008; Buée et al., 2011). Here, more fungal species were found to be related to deciduous trees; however, it is worth mentioning that the proportions of the total volumes of deciduous (65%) and coniferous trees (28%) were biased in our sampling units. The number of fungal species in oak, beech and -dominated stands was similar with respect to the total species pool of wood-inhabiting fungi (Supplementary Fig 1). The relatively strong effect of tree species on the community composition of wood-inhabiting fungi could be due to the great compositional diversity of tree species in the region. It is known that wood- inhabiting fungi are mainly substrate restricted, as they live within the wood, and species are often selective for certain tree taxa (Boddy and Heilmann-Clausen, 2008). It was also underlined by other studies that tree species identity has a marked impact on wood-inhabiting

21

fungal species composition across various spatial scales: indirectly at the centimeter scale via the species specific variation of the chemical environment within the wood, along pH

(Schmidt, 2006) and compositional differences of compounds (Boddy, 1992; Renvall, 1995;

Boddy, 2001), and directly at a stand scale (e.g. Heilmann-Clausen et al., 2005; Sippola et al.,

2005; McMullan-Fisher et al., 2009) and at a continental scale along the distribution of major forest types (Heilmann-Clausen and Boddy, 2008).

In this work, terricolous saprotrophic fungal species composition was found to be shaped by tree species composition (Fig 2b). Previous studies have also confirmed this finding by pointing out a positive response to tree species diversity at the stand scale

(McMullan-Fisher et al., 2009; O’Hanlon and Harrington, 2012) or even a negative one

(Ferris et al., 2000). Terricolous saprotrophic fungi are thought to be mainly a substrate restricted functional group (Gebauer and Taylor, 1999; Boddy et al., 2008) and tree species composition may affect them via the fundamental impacts of tree species on litter quality and quantity.

Regarding the EcM fungi, a contrasting response was revealed to tree species composition: NMDS found no significant effects, but RDA highlighted the proportion of beech to have the strongest importance on EcM species composition (Supplementary Fig 6).

However, many previous studies (e.g. Såstad, 1995; Ferris et al., 2000; Kernaghan et al.,

2003; Morris et al., 2009) supported the idea that EcM species composition is determined principally by the species composition of their host trees at stand scale, but many other studies came to contradictory conclusions highlighting soil properties (e.g. Talbot et al., 2013; Suz et al., 2014) or other biotic factors (e.g. Kennedy, 2010; Peay et al., 2010) to have determinative effects. The picture is not clear, because there is usually a striking contrast between the great diversity of EcM fungal communities and the relatively species-poor stands of host trees in temperate forests (Tedersoo et al., 2014). A large number of EcM fungal species can be found

22

on the root surface even of the same tree individual or root tip (Bahram et al., 2011), and until this complexity is better understood at finer scales, results suggesting changes in EcM species composition at a stand scale are a matter of debate (Erland and Taylor, 2003).

Stand structure

In this study, the total cover of FWD and CWD had a significant effect on wood- inhabiting fungal species composition (Fig 2a), but CWD volume alone was not important.

Many studies (reviewed in Lonsdale et al., 2008) detected the quantity of dead wood to have the highest influence on wood-inhabiting fungi. However, the influence of FWD and CWD on wood-inhabiting fungi cannot be separated in this study, but a considerable impact of FWD was revealed. Here, CWD was selective for only a very low proportion (12%) of fungal taxa

(details in Supplementary Table 1). It is probably because out of the total CWD volume on sites, oak and conifer logs in decay stages 2–3 amounted to 44% which is mainly heartwood and hence, the most species-poor CWD type (Boddy and Heilmann-Clausen, 2008).

Comparative studies in Europe (e.g. Küffer et al., 2008; Abrego and Salcedo, 2011) have also suggested that a large proportion of wood-inhabiting fungi can be harboured on FWD in managed forests.

The density of large trees (in NMDS, Fig 2b and c) and the mean DBH of trees (in

RDA, Supplementary Fig 4 and 6) were found to be significant in structuring the species composition of each functional group. However, these two variables were moderately correlated (r = 0.381, p = 0.024), but both of them may have the same effect on fungal communities influencing them via the presence of large trees in the forest stands. Only the

EcM community was shaped considerably by both of these factors, but such a result, based on sporocarp data, is impossible to interpret adequately. However, large trees can serve as “hubs” in the common mycorrhizal network belowground (reviewed in Simard et al., 2012), or stands

23

in different successional phases (with different tree sizes) can harbour distinctive EcM communities (Smith et al., 2002; Twieg et al., 2007) that can both make changes in sporocarp occurrences.

The relative volume of decayed logs was revealed to have a significant effect on EcM community composition (Fig 2c). It is proven that the majority of EcM fungi evolved from humus and wood saprotrophic ancestors (Tedersoo et al., 2010), therefore many EcM fungi still have some ability to decompose wood in latter decay stages. A similar EcM community response was revealed by Walker et al. (2012) to CWD volume in their clear-cut forest system emphasizing that dead wood provides a balanced environment for fungi with respect to microclimate and available nutrients. By contrast, it was hypothesized by Baldrian (2009) that the lignocellulose-decomposing enzymes of EcM fungi may support only escape from a dying root.

Soil and litter conditions

The pH of litter determined the species composition of terricolous saprotrophic fungi and had a considerable effect on EcM fungi (Fig 2b and c). Similar influences of soil pH have already been published on terricolous saprotrophic (Ferris et al., 2000; Talbot et al., 2013) and

EcM community composition (Baar and ter Braak, 1996). It was shown in the present study that the underlying litter pH gradient with a determinative effect on terricolous saprotrophic community was related to Scots pine proportion highlighting that the tree species composition has a strong impact on litter pH, and Scots pine has a more acidic litter compared to that of deciduous trees (Augusto et al., 2003).

The weak, but significant effects of soil N, P and K contents on terricolous (EcM and terricolous saprotrophic) communities (Fig 2b and c, Supplementary Fig 5a) cannot be explained without mentioning their relatively strong collinearity (|r| = 0.4–0.5) compared to

24

their relations to the ordination axes (|r| = 0.3–0.5). However, similar relationships were detected by Baar and ter Braak (1996) and Toljander et al. (2006) among N, P and K contents who suggested that K likely plays a minor role compared to N and P (nutrients) in the occurrence of fungi. Soil K content was suggested to be important in the osmoregulation and sporocarp formation (Tyler, 1982). Soil P content was reported by Conn and Dighton (2000) and Morris et al. (2009) to have important consequences for EcM community development, but in our study, only a marginal significance of soil P was detected. In Supplementary Fig

5b, the general negative impact of soil N content on terricolous saprotrophic fungi in the ŐNP is given. In other European countries, concordant (e.g. Buée et al., 2011) and contradictory

(e.g. Tarvainen et al., 2003) results were also detected. However, numerous N fertilization experiments have been conducted (e.g. Tarvainen et al., 2003) on terricolous macrofungi, but in most cases the fruiting of EcM communities was negatively affected.

Microclimate

Supported by RDA only, wood-inhabiting and terricolous saprotrophic communities were structured by air temperature (Supplementary Fig 4 and 5). Regarding wood-inhabiting fungi, this result is in agreement with those of Boddy (1992, 2001) within the wood and

Renvall (1995) at a stand scale and Heilmann-Clausen et al. (2014) at a continental scale. The general effect of air temperature on wood-inhabiting fungi is too difficult to interpret in our study. In contrast, air temperature had a clear negative effect on the majority of terricolous saprotrophic fungal species. According to Berg and McClaugherty (2014), the optimal temperature is vital for the right activity of cellulo- and ligninolytic enzymes of this functional group. In accordance, their optimal temperatures in the studied region may be in rather closed stands with no sun-exposed litter layer.

25

Other factors

Revealed by RDA only, management history and landscape characteristics were demonstrated with low and moderate effects on wood-inhabiting and EcM communities, respectively (Supplementary Fig 4 and 6). The negative effects of forest management on wood-inhabiting fungi is widely studied (e.g. Lindner et al., 2006), but such a clear community response was not detected here. Only the EcM community was influenced by landscape characteristics indicating that this group is affected significantly also at larger (r =

300 m) scales compared to the other studied functional groups.

Limitations of data

Our community data is biased by all the disadvantages of using sporocarp incidences to estimate macrofungal abundance [see Tóth and Barta (2010) for a review]. The biggest weakness is the short duration (2 yr) of our field visits that can only provide an underestimate of fungal species richness in the sampling units. It has been shown that additional species can also be found after 21 years of surveys (Straatsma et al., 2001). Another potential source of error is the variation among years in the fruiting of species (Fernández-Toirán et al., 2006), which also was observed in this study. Given these limitations, the current results must therefore be viewed with caution.

Conclusions

It is hypothesized that substrate properties, tree species composition and microclimate, in that order, are the most influential drivers of fungal species composition in the studied region, and their relative influences differ among functional groups. Wood-inhabiting fungal species composition was driven primarily by the species composition of living trees, while substrate properties and microclimate had minor relevance. The terricolous saprotrophic

26

community was determined principally by a litter pH gradient involving tree species proportions and soil/litter properties. Microclimate had no concordant effect. The EcM fungal species composition was not structured by obvious ecological gradients supported simultaneously by more than one environmental variable, but litter pH and tree size had significant effects. The lack of detected gradients suggests that the most important drivers of

EcM fungi remained unmeasured. Regarding each functional group, no clear responses to management history or to the surrounding landscape were found. However, it was confirmed that macrofungal communities are related significantly to environmental drivers that are relatively easy to measure at a stand scale. To gain a further insight into stand-scale drivers of fungal species composition, sporocarp surveys should be combined with DNA sequence based sampling methods from a below-ground study.

27

Acknowledgements

GK had the main role in this study which is the part of his PhD thesis. Fieldwork and identification were headed by IS; and all authors participated in the sporocarp surveys.

Identifications were completed by IS, BD, GK and KT. AB was responsible for soil and litter data. This study was carried out in the frame of the Őrs-erdő Project

(http://orserdo.okologia.mta.hu/), which was headed by PÓ who provided continuous participation and assistance in data analyses and writing. The study was supported by the

Hungarian Scientific Research Fund (OTKA, K79158) and the Directorate of Őrség National

Park. PÓ was supported by the János Bolyai Research Scholarship of the Hungarian Academy of Sciences. The authors are very grateful to two anonymous reviewers for valuable comments on the manuscript.

28

References

Abrego, N., Salcedo, I., 2011. How does fungal diversity change based on woody debris type?

A case study in Northern Spain. Ekologija 57, 109–119.

Arcanum, 2006. Secondary Military Survey of the Habsburg Empire 1806–1869. DVD-ROM,

Arcanum Kft., Budapest.

Augusto, L., Dupouey, J.-L., Ranger, J., 2003. Effects of tree species on understory vegetation

and environmental conditions in temperate forests. Annals of Forest Science 60,

823–831.

Baar, J., ter Braak, C.J.F., 1996. Ectomycorrhizal sporocarp occurrence as affected by

manipulation of litter and humus layers in Scots pine stands of different age. Applied

Soil Ecology 4, 61–73.

Babos, M., 1989. Magyarország kalaposgombáinak ( s.l.) jegyzéke - I. (The

Agaricales s.l. taxa of Hungary. I.). Mikológiai Közlemények, Clusiana 1–3, 3–234.

Bahram, M., Põlme, S., Kõljalg, U., Tedersoo, L., 2011. A single European aspen (Populus

tremula) tree individual may potentially harbour dozens of Cenococcum geophilum

ITS genotypes and hundreds of species of ectomycorrhizal fungi. FEMS

Microbiology Ecology 75, 313–320.

Baldrian, P., 2009. Ectomycorrhizal fungi and their enzymes in soils: is there enough

evidence for their role as facultative soil saprotrophs? Oecologia 161, 657–660.

Bässler, C., Müller, J., Dziock, F., Brandl, R., 2010. Effects of resource availability and

climate on the diversity of wood-decaying fungi. Journal of Ecology 98, 822–832.

Bellér, P., 1997. Talajvizsgálati módszerek (Methods of soil analysis). Egyetemi jegyzet,

Soproni Egyetem, Erdőmérnöki Kar, Termőhelyismerettani Tanszék, Sopron.

Berg, B., McClaugherty, Ch., 2014. Plant Litter: Decomposition, humus formation, carbon

sequestration, Third Ed, Springer-Verlag, Berlin Heidelberg.

29

Bernicchia, A., Benni, A., Venturella, G., Gargano, M.L., Saitta, A., Gorjón, S.P., 2008.

Aphyllophoraceous wood-inhabiting fungi on Quercus spp. in Italy. Mycotaxon 104,

425–428.

Bernicchia, A., Gorjón, S.P., 2010. Corticiaceae s.l., Fungi Europaei Vol. 12. Edizioni

Candusso, Alassio, pp. 638–639.

Bernicchia, A., Savino, E., Gorjón, S.P., 2007. Aphyllophoraceous wood-inhabiting fungi on

Pinus spp. in Italy. Mycotaxon 101, 5–8.

Bernicchia, A., Venturella, G., Saitta, A., Gorjón, S.P., 2007. Aphyllophoraceous wood-

inhabiting fungi on Fagus sylvatica in Italy. Mycotaxon 101, 229–232.

Boddy, L., 1992. Development and function of fungal communities in decomposing wood. In

Carroll, G.C., Wicklow, D.T. (Eds) The fungal community: Its organization and role

in the ecosystem, Second Ed, Marcel Dekker Inc., New York, pp. 749–782.

Boddy, L., 2001. Fungal community ecology and wood decomposition processes in

angiosperms: from standing tree to complete decay of coarse woody debris.

Ecological Bulletins 49, 43–56.

Boddy, L., Frankland, J.C., van West, P. (Eds), 2008. Ecology of saprotrophic

basidiomycetes. The British Mycological Society, Academic Press, London.

Boddy, L., Heilmann-Clausen, J., 2008. Basidiomycete community development in temperate

angiosperm wood. In Boddy, L., Frankland, J.C., van West, P. (Eds) Ecology of

saprotrophic basidiomycetes. The British Mycological Society, Academic Press,

London, pp. 211–237.

Bohus, G., 1973. Soil acidity and the occurrence of fungi in deciduous forests. Annales

Historico-Naturales Musei Nationalis Hungarici 65, 63–81.

Bohus, G., Babos, M., 1967. Mycocoenological investigation of acidophilous deciduous

forests in Hungary. Botanischer Jahrbücher 87, 304–360.

30

Breitenbach, J., Kränzlin, F., 1986. Fungi of Switzerland, Vol. 2. Mykologia, Luzern, p. 366.

Buée, M., Maurice, J.P., Zeller, B., Andrianarisoa, S., Ranger, J., Courtecuisse, R., Marçais,

B., Le Tacon, F., 2011. Influence of tree species on richness and diversity of

epigeous fungal communities in a French temperate forest stand. Fungal Ecology 4,

22–31.

Büntgen, U., Kauserud, H., Egli, S., 2012. Linking climate variability to

productivity and phenology. Frontiers in Ecology and the Environment 10, 14–19.

Christensen, M., Heilmann-Clausen, J., 2013. The genus Tricholoma, Fungi of Northern

Europe, Vol. 4. Narayana Press, Gylling, pp. 194–195.

Ciortan, I., 2009. Contributions to the knowledge diversity of lignicolous macromycetes

(Basidiomycetes) from Căpățânii Mountains. The Annals of Oradea University,

Biology Fascicle 16, 53–59.

Claridge, A.W., Barry, S.C., Cork, S.J., Trappe, J.M., 2000. Diversity and habitat

relationships of hypogeous fungi II. Factors influencing the occurance and number of

taxa. Biodiversity and Conservation 9, 175–199.

Conn, Ch., Dighton, J., 2000. Litter quality influences on decomposition, ectomycorrhizal

community structure and mycorrhizal root surface acid phosphatase activity. Soil

Biology & Biochemistry 32, 489–496.

Courty, P.-E., Franc, A., Pierrat, J.-C., Garbaye, J., 2008. Temporal changes in the

ectomycorrhizal community in two soil horizons of a temperate oak forest. Applied

and Environmental Microbiology 74, 5792–5801.

Cox, F., Barsoum, N., Lilleskov, E.A., Bidartondo, M.I., 2010. Nitrogen availability is a

primary determinant of conifer mycorrhizas across complex environmental gradients.

Ecology Letters 13, 1103–1113.

Dövényi, Z., 2010. Cadastre of Hungarian Regions. MTA Földrajztudományi Intézet,

31

Budapest.

Erland, S., Taylor, A.F.S., 2003. Diversity of ecto-mycorrhizal fungal communities in relation

to the abiotic environment. In van der Heijden, M.G.A., Sanders, I.R. (Eds),

Mycorrhizal Ecology, Second Ed, Ecological Studies, Vol. 157, Springer-Verlag,

Berlin, Heidelberg, pp. 163–200.

Fernández-Toirán, L.M., Ágreda, T., Olano, J.M., 2006. Stand age and sampling year effect

on the fungal fruit body community in Pinus pinaster forests in central Spain.

Canadian Journal of Botany 84, 1249–1258.

Ferris, R., Peace, A.J., Newton, A.C., 2000. Macrofungal communities of lowland Scots pine

(Pinus sylvestris L.) and Norway spruce [Picea abies (L.) Karsten.] plantations in

England: relationships with site factors and stand structure. Forest Ecology and

Management 131, 255–267.

Fischer, M., Wagner, T., 1999. RFLP analysis as a tool for identification of lignicolous

basidiomycetes: European polypores. European Journal of Forest Pathology 29, 295–

304.

Fox, J., 2005. The R Commander: A Basic Statistics Graphical User Interface to R. Journal of

Statistical Software 14(9), 1–42.

Galli, R., 2006. I Lattari. dalla Natura, Milano, pp. 56, 218, 222.

Gebauer, G., Taylor, A.F.S., 1999. 15N natural abundance in fruit bodies of different

functional groups of fungi in relation to substrate utilization. New Phytologist 142,

93–101.

Gómez-Hernández, M., Williams-Linera, G., Guevara, R., Lodge, D.J., 2012. Patterns of

macromycete community assemblage along an elevation gradient: options for fungal

gradient and metacommunity analyses. Biodiversity and Conservation 21, 2247–

2268.

32

Gyöngyössy, P., 2008. “Gyántásország”: Történeti adatok az őrségi erdők erdészeti és

természetvédelmi értékeléséhez (“Gyántásország”: Evaluation of forests in Őrség in

forestry and conservation perspectives using historical data). Ciklámen Füzetek,

Kerekerdő Alapítvány, Szombathely.

Halász, G., (Ed.) 2006. Forest regions of Hungary. Hungarian State Forest Service, Budapest.

Halme, P., Ódor, P., Christensen, M., Piltaver, A., Veerkamp, M., Walleyn, R., Siller, I.,

Heilmann-Clausen, J., 2013. The effects of habitat degradation on metacommunity

structure of wood-inhabiting fungi in European beech forests. Biological

Conservation 168, 24–30.

Heilmann-Clausen, J., 2001. A gradient analysis of communities of macrofungi and slime

moulds on decaying beech logs. Mycological Research 105, 575–596.

Heilmann-Clausen, J., 2005. Diversity of saproxylic fungi on decaying beech wood in

protected forests in the county of Halland. Scientific Report, HabitatVision, pp. 64.

Heilmann-Clausen, J., Aude, E., Christensen, M., 2005. Cryptogam communities on decaying

deciduous wood – does tree species diversity matter? Biodiversity and Conservation

14, 2061–2078.

Heilmann-Clausen, J., Aude, E., van Dort, K., Christensen, M., Piltaver, A., Veerkamp, M.,

Walleyn, R., Siller, I., Standovár, T., Ódor, P., 2014. Communities of wood-

inhabiting bryophytes and fungi on dead beech logs in Europe – reflecting substrate

quality or shaped by climate and forest conditions? Journal of Biogeography 41,

2269–2282.

Heilmann-Clausen, J., Boddy, L., 2008. Distribution patterns of wood-decay basidiomycetes

at the landscape to global scale. In Boddy, L., Frankland, J.C., van West, P. (Eds)

Ecology of saprotrophic basidiomycetes. The British Mycological Society, Academic

Press, London, pp. 263–275.

33

Heilmann-Clausen, J., Christensen, M., 2003a. Diversity patterns and community structure of

wood-inhabiting macrofungi on beech in the Danish landscape. In Heilmann-

Clausen, J., Wood-inhabiting fungi in Danish deciduous forests – diversity, habitat

preferences and conservation. PhD dissertation, Royal Veterinary and Agricultural

University, Frederiksberg, Denmark.

Heilmann-Clausen, J., Christensen, M., 2003b. Fungal diversity on decaying beech logs –

implications for sustainable forestry. Biodiversity and Conservation 12, 953–973.

Heilmann-Clausen, J., Christensen, M., 2004. Does size matter? On the importance of various

dead wood fractions for fungal diversity in Danish beech forests. Forest Ecology and

Management 201, 105–117.

Humphrey, J.W., Newton, A.C., Peace, A.J., Holden, E., 2000. The importance of conifer

plantations in northern Britain as a habitat for native fungi. Biological Conservation

96, 241–252.

ISO 10694, 1995. Soil Quality – Determination of organic and total carbon after dry

combustion (elementary analysis). International Organization for Standardization,

Geneva. p. 7. (www.iso.org)

ISO 13878, 1998. Soil Quality – Determination of total nitrogen content by dry combustion

(“elemental analysis”). International Organization for Standardization, Geneva. p. 5.

(www.iso.org)

Jones, M.D., Durall, D.M., Cairney, J.W.G., 2003. Ectomycorrhizal fungal communities in

young forest stands regenerating after clearcut logging. New Phytologist 157, 399–

422.

Juhász, P., Bidló, A., Heil, B., Kovács, G., Ódor, P., 2011. Őrségi erdőtalajok széntartalmi

vizsgálata (Investigation of carbon content of forest soils in Őrség, West Hungary).

Talajvédelem (suppl.), Szeged. pp. 377–382.

34

Kacprzyk, M., Bednarz, B. Kuźnik, E., 2014. Dead trees in beech stands of the Bieszczady

National Park: quantitative and qualitative structure of associated macrofungi.

Applied Ecology and Environmental Research 12, 325–344.

Kennedy, P., 2010. Ectomycorrhizal fungi and interspecific competition: species interactions,

community structure, coexistence mechanisms, and future research directions. New

Phytologist 187, 895–910.

Kennedy, P.G., Peay, K.G., Bruns, T.D., 2009. Root tip competition among ectomycorrhizal

fungi: are priority effects a rule or an exception? Ecology 90, 2098–2107.

Kernaghan, G., Widden, P., Bergeron, Y., Légaré, S., Paré, D., 2003. Biotic and abiotic

factors affecting ectomycorrhizal diversity in boreal mixed-woods. Oikos 102, 497–

504.

Knudsen, H., Vesterholt, J. (Eds), 2012. Funga Nordica: Agaricoid, boletoid, clavarioid,

cyphelloid and gastroid genera I–II. Second Ed, Nordsvamp, Copenhagen.

Krieglsteiner, G.J. (Ed), 2000. Die Groβpilze Baden-Württembergs, Band 2. Verlag Eugen

Ulmer GmbH & Co., Stuttgart, pp. 88–89.

Krieglsteiner, G.J. (Ed), 2001. Die Groβpilze Baden-Württembergs, Band 3. Verlag Eugen

Ulmer GmbH & Co., Stuttgart, pp. 133–134, 432–433.

Kruskal, J.B., 1964. Multidimensional scaling by optimizing goodness of fit to a nonmetric

hypothesis. Psychometrika 29, 1–27.

Küffer, N., Gillet, F., Senn-Irlet, B., Aragno, M., Job, D., 2008. Ecological determinants of

fungal diversity on dead wood in European forests. Fungal Diversity 30, 83–95.

Lakatos, F., Molnár, M., 2009. Mass mortality of beech (Fagus sylvatica L.) in South-West

Hungary. Acta Silvatica et Lignaria Hungarica 5, 75–82.

Lang, C., Seven, J., Polle, A., 2011. Host preferences and differential contributions of

deciduous tree species shape mycorrhizal species richness in a mixed Central

35

European forest. Mycorrhiza 21, 297–308.

Legendre, P., Legendre, L., 1998. Numerical Ecology, Second Ed, Elsevier Science B.V.,

Amsterdam.

Lepš, J., Šmilauer, P., 2003. Multivariate analysis of ecological data using Canoco.

Cambridge University Press, New York.

Lilleskov, E.A., Parrent, J.L., 2007. Can we develop general predictive models of mycorrhizal

fungal community–environment relationships? New Phytologist 174, 250–256.

Lindblad, I., 2001. Diversity of poroid and some corticoid wood-inhabiting fungi along the

rainfall gradient in tropical forests, Costa Rica. Journal of Tropical Ecology 17, 353–

369.

Lindner, D.L., Burdsall, H.H., Stanosz, G.R., 2006. Species diversity of polyporoid and

in northern hardwood forests with differing management histories.

Mycologia 98, 195–217.

Lonsdale, D., Pautasso, M., Holdenrieder, O., 2008. Wood-decaying fungi in the forest:

conservation needs and management options. European Journal of Forest Research

127, 1–22.

McCune, B., Grace, J.B., 2002. Analysis of ecological communities. MjM Software Design,

Oregon.

McMullan-Fisher, S.J.M., 2008. Surrogates for cryptogam conservation – associations

between mosses, macrofungi, vascular plants and environmental variables. PhD

dissertation, University of Tasmania, School of Geography and Environmental

Studies, Hobart.

McMullan-Fisher, S.J.M., Kirkpatrick, J.B., May, T.W., Pharo, E.J., 2009. Surrogates for

macrofungi and mosses in reservation planning. Conservation Biology 24, 730–736.

Miettinen, O., Niemelä, T., Spirin, W., 2006. Northern Antrodiella species: the identity of A.

36

semisupina, and type studies of related taxa. Mycotaxon 96, 211–239.

Morris, M.H., Pérez-Pérez, M.A., Smith, M.E., Bledsoe, C.S., 2009. Influence of host species

on ectomycorrhizal communities associated with two co-occurring oaks (Quercus

spp.) in a tropical cloud forest. FEMS Microbiology Ecology 69, 274–287.

O’Hanlon, R., Harrington, T.J., 2012. Macrofungal diversity and ecology in four Irish forest

types. Fungal Ecology 5, 499–508.

Ódor P., Heilmann-Clausen, J., Christensen, M., Aude, E., van Dort, K.W., Piltaver, A.,

Siller, I., Veerkamp, M.T., Walleyn, R., Standovár, T., van Hees, A.F.M., Kosec, J.,

Matočec, N., Kraigher, H., Grebenc, T., 2006. Diversity of dead wood inhabiting

fungi and bryophytes in semi-natural beech forests in Europe. Biological

Conservation 131, 58–71.

Ódor, P., van Hees, A.F.M., 2004. Preferences of dead wood inhabiting bryophytes for decay

stage, log size and habitat types in Hungarian beech forests. Journal of Bryology 26,

79–95.

Oksanen, J., 2013. Multivariate analysis of ecological communities in R: vegan tutorial.

(http://cc.oulu.fi/~jarioksa/opetus/metodi/vegantutor.pdf)

Oksanen, J., Blanchet, F.G., Kindt, R., Legendre, P., Minchin, P.R., O'Hara, R.B., Simpson,

G.L., Solymos, P., Stevens, M.H.H., Wagner, H., 2013. Vegan: Community Ecology

Package. R package version 2.0-8. (http://CRAN.R-project.org/package=vegan)

Őrs-erdő Project, 2015. The effect of stand structure on the composition and diversity of

different organism groups in Őrség (Western Hungary).

(http://orserdo.okologia.mta.hu/)

Pál-Fám, F., 2001. A Mecsek hegység nagygombái (The macrofungi of the Mecsek Mts.).

Mikológiai Közlemények, Clusiana 40, 5–66.

Papp, V., 2013. Corticioid basidiomycetes of Hungary I. The genus Hymenochaete.

37

Mikológiai Közlemények, Clusiana 52, 45–56.

Peay, K.G., Garbelotto, M., Bruns, T.D., 2010. Evidence of dispersal limitation in soil

microorganisms: isolation reduces species richness on mycorrhizal tree islands.

Ecology 91, 3631–3640.

Prentice, I.C., 1977. Non-metric ordination methods in ecology. Journal of Ecology 65, 85–

94.

R Core Team, 2013. R: a language and environment for statistical computing. R Foundation

for Statistical Computing, Vienna, Austria. (http://www.R-project.org)

Renvall, P., 1995. Community structure and dynamics of wood-rotting Basidiomycetes on

decomposing conifer trunks in northern Finnland. Karstenia 35, 1–51.

Reverchon, F., del Ortega-Larrocea, P.M., Pérez-Moreno, J., 2010. Saprophytic fungal

communities change in diversity and species composition across a volcanic soil

chronosequence at Sierra del Chichinautzin, Mexico. Annals of Microbiology 60,

217–226.

Rimóczi, I., Jeppson, M., Benedek, L., 2011. Characteristic and rare species of

Gasteromycetes in Eupannonicum. Fungi non Delineati, Pars 56–57, Edizioni

Candusso, Alassio, pp. 105–108.

Ryvarden, L., Gilbertson, R.L., 1994. European polypores (Meripilus–Tyromyces), Part 2,

Fungiflora, Oslo, pp. 667–669.

Salerni, E., Laganà, A., Perini, C., Loppi, S., De Domonicus, V., 2002. Effects of temperature

and rainfall on fruiting of macrofungi in oak forests of the Mediterranean area. Israel

Journal of Plant Sciences 50, 189–198.

Såstad, S.M., 1995. Fungi – vegetation relationships in a Pinus sylvestris forest in central

Norway. Canadian Journal of Botany 73, 807–816.

Sato, H., Morimoto, S., Hattori, T., 2012. A thirty-year survey reveals that ecosystem function

38

of fungi predicts phenology of mushroom fruiting. PLoS ONE 7, e49777.

Schmidt, O., 2006. Wood and tree fungi: biology, damage, protection, and use. Springer-

Verlag, Berlin.

Shannon, C.E., Weaver, W., 1949. The mathematical theory of communication. University of

Illinois Press, Urbana.

Sibson, R., 1972. Order invariant methods for data analysis. Journal of the Royal Statistical

Society, Series B (Methodological) 34, 311–349.

Siitonen, J., 2001. Forest management, coarse woody debris and saproxylic organisms:

Fennoscandian boreal forests as an example. Ecological Bulletins 49, 11–41.

Siller, I., 2004. Hazai montán bükkös erdőrezervátumok (Mátra: Kékes Észak, Bükk: Őserdő)

nagygombái [Macrofungi of montane beech forest reserves (Mátra Mountains: Kékes

Észak, Bükk Mountains: Őserdő)]. PhD dissertation, Budapesti

Közgazdaságtudományi és Államigazgatási Egyetem, Budapest.

Siller, I., Kutszegi, G., Takács, K., Varga, T., Merényi, Zs., Turcsányi, G., Ódor, P., Dima, B.,

2013. Sixty-one macrofungi species new to Hungary in Őrség National Park.

Mycosphere 4, 871–924.

Simard, S.W., Beiler, K.J., Bingham, M.A., Deslippe, J.R., Philip, L.J., Teste, F.P., 2012.

Mycorrhizal networks: Mechanisms, ecology and modelling. Fungal Biology

Reviews 26, 39–60.

Sippola, A.-L., Mönkkönen, M., Renvall, P., 2005. Polypore diversity in the herb-rich

woodland key habitats of Koli National Park in eastern Finland. Biological

Conservation 126, 260–269.

Smith, J.E., Molina, R., Huso, M.M.P., Luoma, D.L., McKay, D., Castellano, M.A., Lebel, T.,

Valachovic, Y., 2002. Species richness, abundance, and composition of hypogeous

and epigeous ectomycorrhizal fungal sporocarps in young, rotation-age, and old-

39

growth stands of Douglas-fir (Pseudotsuga menziesii) in the Cascade Range of

Oregon, U.S.A. Canadian Journal of Botany 80, 186–204.

Smith, S.E., Read, D.J., 2008. Mycorrhizal symbiosis, Third Ed, Elsevier Ltd., Oxford.

Sopp, L., Kolozs, L., 2000. Fatömegszámítási táblázatok (Volume tables for tree species).

Állami Erdészeti Szolgálat, Budapest.

Straatsma, G., Ayer, F., Egli, S., 2001. Species richness, abundance, and phenology of fungal

fruit bodies over 21 years in a Swiss forest plot. Mycological Research 105, 515–

523.

Sundqvist, M.K., Sanders, N.J., Wardle, D.A., 2013. Community and ecosystem responses to

elevational gradients: Processes, mechanisms, and insights for global change. Annual

Review of Ecology, Evolution, and Systematics 44, 261–280.

Suz, L.M., Barsoum, N., Benham, S., Dietrich, H.P., Fetzer, K.D., Fischer, R., García, P.,

Gehrman, J., Kristöfel, F., Manninger, M., Neagu, S., Nicolas, M., Oldenburger, J.,

Raspe, S., Sánchez, G., Schröck, H.W., Schubert, A., Verheyen, K., Verstraeten, A.,

Bidartondo, M.I., 2014. Environmental drivers of ectomycorrhizal communities in

Europe’s temperate oak forests. Molecular Ecology 23, 5628–5644.

Szabó, I., 2012. Poroid fungi of Hungary in the collection of Zoltán Igmándy. Acta Silvatica

et Lignaria Hungarica 8, 113–122.

Szemere, L., 1955. A magyarországi Inocybe-fajok, tekintettel az európai fajokra (The

Inocybe species of Hungary). Annales Historico-Naturales Musei Nationalis

Hungarici 47, 121–154.

Talbot, J.M., Bruns, T.D., Smith, D.P., Branco, S., Glassman, S.I., Erlandson, S., Vilgalys, R.,

Peay, K.G., 2013. Independent roles of ectomycorrhizal and saprotrophic

communities in soil organic matter decomposition. Soil Biology & Biochemistry 57,

282–291.

40

Tarvainen, O., Markkola, A.M., Strömmer, R., 2003. Diversity of macrofungi and plants in

Scots pine forests along an urban pollution gradient. Basic and Applied Ecology 4,

547–556.

Tedersoo, L., Bahram, M., Põlme, S., Kõljalg, U., Yorou, N. S., Wijesundera, R., Ruiz, L. V.,

Vasco-Palacios, A. M., Thu, P. Q., Suija, A., Smith, M. E., Sharp, C., Saluveer, E.,

Saitta, A., Rosas, M., Riit, T., Ratkowsky, D., Pritsch, K., Põldmaa, K., Piepenbring,

M., Phosri, Ch., Peterson, M., Parts, K., Pärtel, K., Otsing, E., Nouhra, E.,

Njouonkou, A. L., Nilsson, R. H., Morgado, L. N., Mayor, J., May, T. W.,

Majuakim, L., Lodge, D. G., Lee, S. S., Larsson, K.-H., Kohout, P., Hosaka, K.,

Hiiesalu, I., Henkel, T. W., Harend, H., Guo, L.-D., Greslebin, A., Grelet, G., Geml,

J., Gates, G., Dunstan, W., Dunk, Ch., Drenkhan, R., Dearnaley, J., De Kesel, A.,

Dang, T., Chen, X., Buegger, F., Brearley, F. Q., Bonito, G., Anslan, S., Abell, S.,

Abarenkov, K., 2014. Global diversity and geography of soil fungi. Science 346,

1256688.

Tedersoo, L., May, T.W., Smith, M.E., 2010. Ectomycorrhizal lifestyle in fungi: global

diversity, distribution, and evolution of phylogenetic lineages. Mycorrhiza 20, 217–

263. ter Braak, C.J.F., Šmilauer, P., 2002. Canoco reference manual and CanoDraw for Windows

User’s guide: Software for canonical community ordination (v. 4.5). Microcomputer

Power, Ithaka NY, USA.

Tímár, G., Ódor, P., Bodonczi, L., 2002. The characteristics of forest vegetation of the Őrség

Landscape Protected Area. Kanitzia 10, 109–136.

Tinya, F., Mihók, B., Márialigeti, S., Mag, Zs., Ódor, P., 2009. A comparison of three indirect

methods for estimating understory light at different spatial scales in temperate mixed

forests. Community Ecology 10, 81–90.

41

Toljander, J.F., Eberhardt, U., Toljander, Y.K., Paul, L.R., Taylor, A.F.S., 2006. Species

composition of an ectomycorrhizal fungal community along a local nutrient gradient

in a boreal forest. New Phytologist 170, 873–884.

Tóth, B.B., Barta, Z., 2010. Ecological studies of ectomycorrhizal fungi: an analysis of survey

methods. Fungal Diversity 45, 3–19.

Twieg, B.D., Durall, D.M., Simard, S.W., 2007. Ectomycorrhizal fungal succession in mixed

temperate forests. New Phytologist 176, 437–447.

Tyler, G., 1982. Accumulation and exclusion of metals in peronata and Amanita

rubescens. Transactions of the British Mycological Society 79, 239–245.

Tyler, G., 1991. Effects of litter treatments on the sporophore production of beech forest

macrofungi. Mycological Research 95, 1137–1139.

Tyler, G., 1992. Tree species affinity of decomposer and ectomycorrhizal macrofungi in

beech (Fagus sylvatica L.), oak (Quercus robur L.) and hornbeam (Carpinus betulus

L.) forests. Forest Ecology and Management 47, 269–284. van der Wal, A., Geydan, T.D., Kuyper, T.W., de Boer, W., 2013. A thready affair: Linking

fungal diversity and community dynamics to terrestrial decomposition processes.

FEMS Microbiology Reviews 37, 477–494.

Walker, J.K.M., Ward, V., Paterson, C., Jones, M.D., 2012. Coarse woody debris retention in

subalpine clearcuts affects ectomycorrhizal root tip community structure within

fifteen years of harvest. Applied Soil Ecology 60, 5–15.

Whalley, A.J.S., 1985. The : some ecological considerations. Sydowia, Annales

Mycologici Series II. 38, 369–382.

Winterhoff, W. (Ed), 1992. Fungi in vegetation science. Kluwer Academic Press, The

Netherlands.

42

Table 1. List of the potential environmental variables influencing the species composition of macrofungal communities.

Transfor- Environmental variable Unit Mean (range) mation TREE SPECIES COMPOSITION Species richness of trees number of 5.63 (2–10) ln species/1600 m2 Shannon diversity of tree species – 0.847 (0.097–1.802) ln Relative volume of beech % 27.9 (0.0–94.4) ln Relative volume of hornbeam % 3.9 (0.0–21.8) ln Relative volume of oaks % 36.4 (1.1–98.0) ln Relative volume of Scots pine % 26.2 (0.0–76.9) ln Relative volume of non-dominant trees % 0.02 (0.00–0.17) ln STAND STRUCTURE Density of trees (Diameter at Breast stems/ha 593.39 (217.75– – Height, DBH > 5 cm) 1392.75) Density of large (DBH > 50 cm) trees stems/ha 17.14 (0.00–56.25) ln Density of shrubs and saplings (DBH = 0– stems/ha 952.14 (0.00– ln 5 cm) 4706.25) Basal area of trees m2/ha 32.87 (21.49–42.26) – Mean DBH of trees cm 26.65 (13.70–40.75) – Coefficient of variation of DBH of trees – 0.480 (0.172–0.983) – (DBH > 5 cm) Volume of snags (d > 10 cm) m3/ha 8.99 (0.90–65.02) ln Volume of logs (d > 10 cm) m3/ha 10.51 (0.17–59.48) ln Total volume of logs and snags (d > 10 m3/ha 19.50 (1.93–73.37) ln cm) Relative volume of logs (d > 10 cm) in % 54.86 (8.25–98.61) – decay stages 3–6 Total cover of FWD and CWD m2/ha 261.57 (79.44– ln 729.99) Cover of understory vegetation m2/ha 740.80 (19.19– ln 4829.30) Cover of bryophytes m2/ha 247.37 (16.57– ln 2201.59) SOIL AND LITTER Cover of soil m2/ha 146.75 (8.56–472.22) – Cover of litter m2/ha 9367 (7815–9834) – pH of litter – 5.29 (4.86–5.68) – pH of soil * – 4.33 (3.96–4.84) – Dry litter mass g/900 cm2 147.66 (105.41– – 243.08) Mass proportion of deciduous litter % 14.71 (2.54–32.80) – Mass proportion of decayed litter % 67.71 (51.58–84.16) – Hydrolytic acidity of soil (y1) * – 30.21 (20.68–45.22) – 43

Exchangeable acidity of soil (y2) * – 15.27 (3.94–30.47) – Fine texture (clay and silt) proportion of % 51.95 (27.60–68.60) – soil * Carbon (C) content of litter % 65.69 (42.87–78.09) – Carbon content of soil * % 6.45 (3.30–11.54) – Nitrogen (N) content of litter % 1.28 (0.83–1.84) – Nitrogen content of soil * % 0.22 (0.11–0.34) – Phosphorus (P) content of soil * mg P2O5/100 4.29 (1.96–9.35) – g Potassium (K) content of soil * mg K2O/100 7.74 (4.00–13.10) – g MICROCLIMATE Mean daily air temperature difference °C –0.10 (–0.93–0.73) – Daily air temperature range difference °C 0.94 (–0.42–2.49) – Mean daily air humidity difference % 0.84 (–1.83–3.32) – Daily air humidity range difference % 1.89 (–2.27–6.58) – Mean relative diffuse light % 2.93 (0.62–10.36) ln Coefficient of variation of relative diffuse % 0.51 (0.12–1.23) ln light LANDSCAPE (radius = 300 m) Proportion of cutting areas % 5.73 (0.00–23.03) ln Proportion of forests % 89.80 (56.92–100.00) – Proportion of open patches (settlements, % 4.72 (0.00–45.25) – meadows, arable lands) Shannon diversity of landscape elements – 1.114 (0.108–1.858) – MANAGEMENT HISTORY Historical proportion of forests ** % 76.58 (24.03–100.00) – Historical proportion of meadows ** % 7.26 (0.00–40.73) – Historical proportion of arable lands ** % 16.16 (0.00–61.27) – Locality of forests in 1853 binary 0.800 (0–1) – Locality of arable lands in 1853 binary 0.171 (0–1) – * soil layer: 0–10 cm, ** radius = 300 m

44

Table 2. Species richness and proportions of functional groups.

Terricolous Wood- saprotrophic EcM fungi Other fungi * Totals inhabiting fungi fungi Number of 245 (118) 127 (47) 290 (34) 25 (11) 687 (196) obtained species (genera) Proportion of 36 18 42 4 100 functional groups (%) Descriptive 40.14 18.31 41.17 2.74 102.40 (35.12, statistics of (13.33, 20–83) (11.65, 0–47) (17.13, 14–92) (2.17, 0–7) 38–178) species richness [mean, (SD **, range)] The five most Exidia Auriscalpium Clavulina – – frequent taxa nigricans, vulgare, coralloides, Laccaria Schizopora Gymnopus amethystina, L. flavipora, Sc. peronatus, laccata, Lactarius paradoxa s.l., Leotia lubrica, subdulcis, Russula Stereum Lycoperdon cyanoxantha hirsutum, St. perlatum, ochraceoflavum Mycena pura The richest Mycena (14), Mycena (25), Cortinarius (100), – – genera (number Pluteus (11), (8), Russula (44), of taxa) (8), Gymnopus (8), Inocybe (28), Postia (7) Lyophyllum (6) Lactarius (26) The number of 74 (30%) 35 (28%) 109 (38%) 11 (44%) 229 (33%) species found in one sampling unit * five functional groups involved, ** standard deviation

45

Table 3. Explanatory powers of the variables fitted significantly onto the NMDS results of functional groups (Fig 2a–c). The r2-values are the squared correlation coefficients of the linear regression models built by using the NMDS results as response variables and including each of the environmental variables separately. P-values are based on 999 random permutations of NMDS data.

Environmental variable r2 p-value

WOOD-INHABITING FUNGI Tree species (species richness of trees) 0.3792 0.003 Oaks (relative volume of oaks) 0.3309 0.006 Hydrolytic acidity (hydrolytic acidity of the soil) 0.3235 0.010 Conifers (relative volume of coniferous trees) 0.3172 0.009 Beech (relative volume of beech) 0.2552 0.030 Dead wood (total cover of FWD and CWD) 0.2442 0.035 TERRICOLOUS SAPROTROPHIC FUNGI Scots pine (relative volume of Scots pine) 0.4395 0.002 pH of litter 0.3782 0.005 Soil K (potassium content of the soil) 0.3064 0.026 Density of large (DBH > 50 cm) trees 0.2751 0.043 EcM FUNGI Density of large (DBH > 50 cm) trees 0.3651 0.002 Decayed logs (relative volume of logs in decay stages 3–6) 0.2984 0.010 pH of litter 0.2329 0.039 Phosphorus content of the soil 0.2193 0.056

46

Table 4. Macrofungal taxa with concordant responses to the environment according to both the NMDS and RDA models. Studies (from Central Europe) examining the environmental requirements of fungal species within the European temperate forests are listed. Factors in

Column 2 are detailed in Table 3; the direction of their effect (increasing ↑ or decreasing ↓ units) is depicted.

Influential environmental factors revealed Macrofungal taxa Reference in present study in other sudies

WOOD-INHABITING FUNGI Hypoxylon fragiforme beech↑ beech↑ Kacprzyk et al., 2014 Mycetinis alliaceus beech↑ beech↑ Heilmann-Clausen, 2005 Polyporus varius beech↑ beech↑ Ciortan, 2009 Skeletocutis nivea beech↑ beech↑ Fischer and Wagner, 1999 Xylaria carpophila beech↑ beech (cupule litter)↑ Whalley, 1985 Antrodiella fragrans beech↑ deciduous trees↑ Miettinen et al., 2006 Postia subcaesia beech↑ deciduous trees↑ Siller, 2004; Szabó, 2012 Biscogniauxia beech↑, litter pH↑, mean beech↑ Lakatos and Molnár, 2009 nummularia DBH↑ Xylaria hypoxylon beech↑, litter pH↑, mean beech↑ Heilmann-Clausen, 2005 DBH↑ Trametes versicolor beech↑, litter pH↑, mean deciduous trees↑, Ryvarden and Gilbertson, DBH↑ conifers↓ 1994 Hymenochaete oaks↑, air temperature↑, oaks↑, beech↓, Papp, 2013 rubiginosa hydrolytic acidity↑ hornbeam↓ Schizopora paradoxa oaks↑, air temperature↑, oaks↑, deciduous trees↑ Bernicchia et al., 2007, 2008 s.l. hydrolytic acidity↑ Stereum subtomentosum oaks↑, air temperature↑, oaks↑, deciduous trees↑ Bernicchia et al., 2008 hydrolytic acidity↑ Stereum oaks↑, hydrolytic acidity↑ oaks↑ Bernicchia and Gorjón, 2010 ochraceoflavum Mycena epipterygia pine↑, dead wood↑, air conifers↑ Krieglsteiner, 2001 temperature↓ Ramaria stricta pine↑, dead wood↑, tree mixed (deciduous– Breitenbach and Kränzlin, species↑, air temperature↓ coniferous) stands↑ 1986; Krieglsteiner, 2000 TERRICOLOUS SAPROTROPHIC FUNGI Auriscalpium vulgare pine↑, litter pH↓, density pine (cones)↑ Bernicchia et al., 2007 of large trees↓ Baeospora myosura pine↑, litter pH↓, density conifers↑ Krieglsteiner, 2001 of large trees↓ Lycoperdon molle pine↑, litter pH↓, density mixed stands↑, open Rimóczi et al., 2011 of large trees↓ areas↑, soil pH↑ EcM FUNGI Lactarius blennius beech↑, decayed logs↑ beech↑ Tyler, 1992; Galli, 2006; Lang et al., 2011 Tricholoma ustale beech↑, decayed logs↑ beech↑, deciduous Bohus, 1973; Buée et al., 47

trees↑, soil pH↓ 2011 Russula emetica beech↑, decayed logs↑ beech↑, soil pH↓ Bohus, 1973 Pseudocraterellus beech↑, light↓ hornbeam↑, oaks↑, N Tyler, 1992; Suz et al., 2014 undulatus deposition↑ Inocybe petiginosa beech↑, light↓ oaks↑, soil pH↓ Szemere, 1955; Babos, 1989 Lactarius subdulcis beech↑, soil P↑, litter pH↑ beech↑, soil pH↓ Galli, 2006; Buée et al., 2011 Tricholoma sulphureum beech↑, soil P↓ beech↑, deciduous Christensen and Heilmann- trees↑ Clausen, 2013 Russula undulata soil P↑, light↑, litter pH↓ oaks↑, hornbeam↑, soil Bohus, 1973 pH↓ Lactarius quietus soil P↑, light↑, litter pH↓ oaks↑, soil N↑, soil pH↓ Galli, 2006; Suz et al., 2014 Amanita rubescens soil P↑, light↑, litter pH↓ soil P, N↑, mixed Pál-Fám, 2001; Buée et al., (deciduous–coniferous) 2011 stands↑ Russula nigricans soil P↑, light↑, litter pH↓ soil pH↓, mixed Bohus and Babos, 1967; Pál- (deciduous–coniferous) Fám, 2001 stands↑

48

Figures

Fig 1. Borders of West Hungary; Őrség National Park (440 km2) is highlighted by grey (a).

The geographical positions of the 35 sampling units are indicated by black dots (the underlined sampling units are moderately managed); built-up areas are shown by grey (b). A:

Austria, H: Hungary, HR: Croatia, SK: Slovakia, SLO: Slovenia

49

Fig 2. Local NMDS on wood-inhabiting (a), terricolous saprotrophic (b) and EcM (c) fungal species (black italics) representing the significantly (p < 0.05) fitted environmental variables

(red capitals) and a tri-plot of sampling units (black circles). The most frequent taxa are displayed; the optimal positions of all recorded species are shown in Supplementary Fig 1–3a and b. See Table 3 for the explanatory powers and statistical reliability of environmental variables and Supplementary Table 1 for abbreviations of species. Three dimensional diagrams were plotted; spinning diagrams provide real 3-D views in Supplementary Fig 7–9.

NMDS was run on Bray–Curtis distances. The final stress values, following Kruskal (1964), are multiplied by 100 and were 15.282, 14.331 and 12.186, respectively.

50

51

52

Supplementary Material for the paper entitled “Drivers of macrofungal species composition in temperate forests, West Hungary: functional groups compared”.

Contents

Supplementary Table 1 ...... p. 2 Supplementary Fig 1a–e ...... p. 15 Supplementary Fig 2a–e ...... p. 17 Supplementary Fig 3a–e ...... p. 19 Supplementary Fig 4 ...... p. 21 Supplementary Fig 5a–b ...... p. 22 Supplementary Fig 6 ...... p. 23

53

Supplementary Table 1. List of 687 macrofungi taxa collected and identified in this work. “Code” was generated by using the first three letters of the genus and species name of taxa. Occasionally, other letters were applied to avoid making redundant abbreviations. Six-letter codes are shown for analyzed taxa only. The column “Analyzed by NMDS?” shows the status of macrofungi taxa in the NMDS models: the species indicated by “yes, plotted” (the most abundant ones) are represented in the NMDS diagrams of the printed paper (the remaining taxa highlighted by “yes” are plotted also, but in Supplementary Fig 1–3); the only “omitted” species, Entoloma jahnii, was excluded from NMDS because it was collected exclusively in one of those four omitted (extremely species poor) sampling units that had outlying points in the NMDS of terricolous saprotrophic fungi; “skipped” taxa were excluded due to their currently indefinite trophic status or belonging to a functional group with too low species number for statistical computing. Data are sorted primarily by “Functional group” then by the “Number of occupied sampling units”. Taxa found in less than four sampling units were excluded from the RDAs. Underlined wood-inhabiting taxa were selective for (d > 20 cm) CWD. This list and the description of sampling units are more detailed in: Siller, I., Kutszegi, G., Takács, K., Varga, T., Merényi, Zs., Turcsányi, G., Ódor, P., Dima, B., 2013. Sixty-one macrofungi species new to Hungary in Őrség National Park. Mycosphere 4, 871–924. EcM = ectomycorrhizal; t. sapr. = terricolous saprotrophic; wood-inh. = wood-inhabiting; entomopath. = entomopathogenic; lign./t. sapr. = lignicolous and/or terricolous saprotrophic; t. sapr./myc. = terricolous saprotrophic and/or mycorrhizal Number of Analyzed Functional occupied Macrofungi taxa Author(s) Code by group sampling NMDS? units Laccaria amethystina Cooke Lacame EcM 34 yes, plotted Russula cyanoxantha (Schaeff.) Fr. Ruscya EcM 34 yes, plotted Clavulina coralloides (L.) J. Schröt. Clacor EcM 28 yes, plotted Laccaria laccata (Scop.) Cooke Laclac EcM 27 yes, plotted Lactarius subdulcis (Pers.) Gray Lacsub EcM 27 yes, plotted Xerocomus pruinatus (Fr. & Hök) Quél. Xerpru EcM 26 yes, plotted Lactarius blennius (Fr.) Fr. Lacble EcM 25 yes, plotted Cortinarius decipiens s.l. (Pers.) Fr. Cordec EcM 24 yes, plotted Russula emetica (Schaeff.) Pers. Ruseme EcM 24 yes, plotted Clavulina cinerea (Bull.) J. Schröt. Clacin EcM 23 yes, plotted Pseudocraterellus undulatus (Pers.) Rauschert Pseund EcM 23 yes, plotted Russula fellea (Fr.) Fr. Rusfel EcM 23 yes, plotted Inocybe petiginosa (Fr.) Gillet Inopet EcM 22 yes, plotted Tricholoma sulphureum (Bull.) P. Kumm. Trisul EcM 21 yes, plotted Russula fragilis Fr. Rusfgl EcM 20 yes, plotted Russula nigricans Fr. Rusnig EcM 20 yes, plotted Russula vesca Fr. Rusves EcM 20 yes, plotted Russula undulata Velen. Rusund EcM 19 yes, plotted Amanita rubescens Pers. Amarub EcM 18 yes, plotted Lactarius chrysorrheus Fr. Lacchr EcM 18 yes, plotted Lactarius camphoratus (Bull.) Fr. Laccam EcM 17 yes, plotted Lactarius quietus (Fr.) Fr. Lacqts EcM 17 yes, plotted Tricholoma saponaceum (Fr.) P. Kumm. Trisap EcM 17 yes, plotted Tricholoma ustale (Fr.) P. Kumm. Triust EcM 17 yes, plotted Humaria hemisphaerica (F.H. Wigg.) Fuckel Humhem EcM 15 yes, plotted Lactarius rostratus Heilm.-Claus. Lacros EcM 15 yes, plotted 54

Amanita citrina (Schaeff.) Pers. Amacit EcM 14 yes, plotted Cortinarius flexipes var. flexipes (Pers.) Fr. Corfle EcM 14 yes, plotted Hydnum rufescens Pers. Hydruf EcM 14 yes, plotted Russula heterophylla (Fr.) Fr. Rushet EcM 14 yes, plotted Cantharellus cibarius Fr. Cancib EcM 13 yes Cortinarius sp.15 Cor_15 EcM 13 yes Cortinarius casimiri (Velen.) Huijsman Corcas EcM 13 yes Cortinarius elatior Fr. Corela EcM 13 yes Lactarius vellereus (Fr.) Fr. Lacvel EcM 13 yes Russula acrifolia Romagn. Rusacr EcM 13 yes Russula grata Britzelm. Rusgra EcM 13 yes Clavulina rugosa (Bull.) J. Schröt. Clarug EcM 12 yes Hebeloma velutipes Bruchet Hebvel EcM 12 yes Russula ochroleuca Pers. Rusoch EcM 12 yes Cortinarius anthracinus (Fr.) Sacc. Corant EcM 11 yes Inocybe assimilata Britzelm. Inoass EcM 11 yes Lactarius serifluus (DC.) Fr. Lacser EcM 11 yes Inocybe geophylla (Fr.) P. Kumm. Inogeo EcM 10 yes Cortinarius cagei Melot Corcag EcM 9 yes Cortinarius flexipes var. flabellus (Fr.) H. Lindstr. & Melot Corflf EcM 9 yes Craterellus cornucopioides (L.) Pers. Cracor EcM 9 yes Hebeloma sordescens Vesterh. Hebsor EcM 9 yes Hydnum repandum L. Hydrep EcM 9 yes Lactarius quieticolor Romagn. Lacqtc EcM 9 yes Russula illota Romagn. Rusill EcM 9 yes Cortinarius trivialis s.l. J.E. Lange Cortri EcM 8 yes Lactarius aurantiacus (Pers.) Gray Lacaur EcM 8 yes Paxillus involutus (Batsch) Fr. Paxinv EcM 8 yes Russula sardonia Fr. Russar EcM 8 yes Tricholoma sciodes (Pers.) C. Martín Trisci EcM 8 yes Amanita phalloides (Fr.) Link Amapha EcM 7 yes Cortinarius diasemospermus var. Lamoure Cordia EcM 7 yes diasemospermus Cortinarius infractus s.l. (Pers.) Fr. Corinf EcM 7 yes Cortinarius rigidipes M.M. Moser Corrig EcM 7 yes Cortinarius torvus (Fr.) Fr. Cortor EcM 7 yes Lactarius acris (Bolton) Gray Lacacr EcM 7 yes Lactarius glaucescens Crossl. Lacgla EcM 7 yes Leccinum pseudoscabrum (Kallenb.) Šutara Lecpse EcM 7 yes Russula densifolia Secr. ex Gillet Rusden EcM 7 yes Scleroderma areolatum Ehrenb. Sclare EcM 7 yes Tricholoma album (Schaeff.) P. Kumm. Trialb EcM 7 yes Xerocomus badius (Fr.) E.-J. Gilbert Xerbad EcM 7 yes Amanita argentea Huijsm. Amaarg EcM 6 yes Amanita excelsa (Fr.) Bertill. Amaexc EcM 6 yes Cortinarius tabularis (Fr.) Fr. Cortab EcM 6 yes Craterellus lutescens (Pers.) Fr. Cralut EcM 6 yes Hygrophorus eburneus (Bull.) Fr. Hygebu EcM 6 yes Inocybe cincinnata (Fr.) Quél. Inocin EcM 6 yes Russula mairei Singer Rusmai EcM 6 yes Xerocomus subtomentosus (L.) Quél. Xersub EcM 6 yes Cortinarius subporphyropus Pilát Corsbp EcM 5 yes Cortinarius venetus (Fr.) Fr. Corven EcM 5 yes Hygrophorus poëtarum R. Heim Hygpoe EcM 5 yes Inocybe asterospora Quél. Inoast EcM 5 yes Inocybe fuscidula Velen. Inofus EcM 5 yes Inocybe lilacina (Peck) Kauffman Inolil EcM 5 yes 55

Lactarius fuliginosus (Fr.) Fr. Lacful EcM 5 yes Lactarius pterosporus Romagn. Lacpte EcM 5 yes Lactarius uvidus (Fr.) Fr. Lacuvi EcM 5 yes Leccinum aurantiacum (Bull.) Gray Lecaur EcM 5 yes Russula amoenolens Romagn. Rusamo EcM 5 yes Russula caerulea Fr. Ruscae EcM 5 yes Russula raoultii Quél. Rusrao EcM 5 yes Boletus edulis Bull. Boledu EcM 4 yes Cortinarius sp.14 Cor_14 EcM 4 yes Cortinarius emunctus Fr. Coremu EcM 4 yes Cortinarius largus Fr. Corlgs EcM 4 yes Cortinarius psammocephalus (Bull.) Fr. Corpsa EcM 4 yes Elaphomyces muricatus Fr. Elamur EcM 4 yes Russula aquosa Leclair Rusaqu EcM 4 yes Tricholoma portentosum (Fr.) Quél. Tripor EcM 4 yes Tylopilus felleus (Bull.) P. Karst. Tylfel EcM 4 yes Amanita fulva (Fr.) Fr. Amaful EcM 3 yes Amanita gemmata (Fr.) Bertill. Amagem EcM 3 yes Chroogomphus rutilus (Schaeff.) O.K. Mill. Chrrut EcM 3 yes Cortinarius sp.08 Cor_08 EcM 3 yes Cortinarius acetosus (Velen.) Melot Corace EcM 3 yes Cortinarius acutus s.l. (Pers.) Fr. Coracu EcM 3 yes Cortinarius emollitoides Bidaud, Moënne-Locc. & Coremo EcM 3 yes Reumaux Cortinarius erubescens M.M. Moser Coreru EcM 3 yes Cortinarius hinnuleus s.l. Fr. Corhin EcM 3 yes Cortinarius luhmannii Münzmay, Saar & B. Corluh EcM 3 yes Oertel Cortinarius nolaneiformis (Velenovský) Dima, Cornol EcM 3 yes Niskanen & Liimat. Cortinarius olivaceofuscus Kühner Coroli EcM 3 yes Cortinarius talus Fr. Cortal EcM 3 yes Cortinarius violaceus (L.) Gray Corvio EcM 3 yes Craterellus tubaeformis (Fr.) Quél. Cratub EcM 3 yes Hebeloma cavipes Huijsman Hebcav EcM 3 yes Hebeloma crustuliniforme (Bull.) Quél. Hebcru EcM 3 yes Hygrophorus persoonii Arnolds Hygper EcM 3 yes Inocybe hirtella Bres. Inohir EcM 3 yes Inocybe praetervisa Quél. Inopra EcM 3 yes Inocybe sindonia (Fr.) P. Karst. Inosin EcM 3 yes Lactarius circellatus Fr. Laccir EcM 3 yes Lactarius necator (Bull.) Pers. Lacnec EcM 3 yes Russula chloroides (Krombh.) Bres. Ruschl EcM 3 yes Russula odorata Romagn. Rusodo EcM 3 yes Russula pectinatoides Peck Ruspec EcM 3 yes Russula puellula Ebbesen, F.H. M?ller & Ruspla EcM 3 yes Jul. Schäff. Russula sanguinea (Bull.) Fr. Russan EcM 3 yes Suillus bovinus (L.) Roussel Suibov EcM 3 yes Thelephora palmata (Scop.) Fr. Thepal EcM 3 yes Amanita muscaria (L.) Lam. Amamus EcM 2 yes Amanita vaginata (Bull.) Lam. Amavag EcM 2 yes Boletus reticulatus Schaeff. Bolret EcM 2 yes Cortinarius sp.07 Cor_07 EcM 2 yes Cortinarius sp.22 Cor_22 EcM 2 yes Cortinarius alboviolaceus (Pers.) Fr. Coralv EcM 2 yes Cortinarius balaustinus Fr. Corbal EcM 2 yes 56

Cortinarius bolaris (Pers.) Fr. Corbol EcM 2 yes Cortinarius calochrous (Pers.) Gray Corcal EcM 2 yes Cortinarius cinnabarinus Fr. Corcib EcM 2 yes Cortinarius callisteus (Fr.) Fr. Corcll EcM 2 yes Cortinarius croceus (Schaeff.) Gray Corcro EcM 2 yes Cortinarius duracinus s.l. Fr. Cordur EcM 2 yes Cortinarius glaucopus (Schaeff.) Gray Corgla EcM 2 yes Cortinarius lepidopus Cooke Corlep EcM 2 yes Cortinarius melleopallens (Fr.) Britzelm. Cormll EcM 2 yes Cortinarius nemorensis s. Saar (Fr.) J.E. Lange Cornem EcM 2 yes Cortinarius orellanus Fr. Corore EcM 2 yes Cortinarius praestigiosus (Fr.) M.M. Moser Corpra EcM 2 yes Cortinarius renidens Fr. Corren EcM 2 yes Cortinarius safranopes Rob. Henry Corsaf EcM 2 yes Cortinarius veregregius Rob. Henry Corver EcM 2 yes Cortinarius vibratilis (Fr.) Fr. Corvib EcM 2 yes Hebeloma birrus (Fr.) Sacc. Hebbir EcM 2 yes Hebeloma hiemale Bres. Hebhie EcM 2 yes Hebeloma radicosum (Bull.) Ricken Hebrad EcM 2 yes Hygrophorus russula (Schaeff.) Kauffman Hygrus EcM 2 yes Inocybe calida Velen. Inocal EcM 2 yes Inocybe cervicolor (Pers.) Quél. Inocer EcM 2 yes Inocybe furfurea Kühner Inofur EcM 2 yes Inocybe jacobi Kühner Inojac EcM 2 yes Inocybe mixtilis (Britzelm.) Sacc. Inomix EcM 2 yes Inocybe nitidiuscula (Britzelm.) Lapl. Inonit EcM 2 yes Inocybe pseudoreducta Stangl & Glowinski Inopse EcM 2 yes Inocybe soluta Velen. Inosol EcM 2 yes Lactarius flexuosus (Pers.) Gray Lacfle EcM 2 yes Lactarius fluens Boud. Lacflu EcM 2 yes Lactarius ruginosus Romagn. Lacrug EcM 2 yes Lactarius torminosus (Schaeff.) Pers. Lactor EcM 2 yes Ramaria fennica var. fennica cf. (P. Karst.) Ricken Ram_fe EcM 2 yes Ramaria flavescens cf. (Schaeff.) R.H. Petersen Ram_fl EcM 2 yes Ramaria fennica var. fumigata (Peck) Schild Ramfvf EcM 2 yes Russula fragrantissima Romagn. Rusfgs EcM 2 yes Russula foetens Pers. Rusfoe EcM 2 yes Russula graveolens Romell Rusgrv EcM 2 yes Russula puellaris Fr. Ruspls EcM 2 yes Scleroderma citrinum Pers. Sclcit EcM 2 yes Sebacina incrustans (Pers.) Tul. & C. Tul. Sebinc EcM 2 yes Sistotrema confluens Pers. Siscon EcM 2 yes Suillus variegatus (Sw.) Kuntze Suivar EcM 2 yes Xerocomus cisalpinus Simonini, H. Ladurner & Xercis EcM 2 yes Peintner Xerocomus ferrugineus (Schaeff.) Alessio Xerfer EcM 2 yes Xerocomus porosporus Imler Xerpor EcM 2 yes Amanita eliae Quél. Amaeli EcM 1 yes Amanita franchetii (Boud.) Fayod Amafra EcM 1 yes Amanita porphyria Alb. & Schwein. Amapor EcM 1 yes Chalciporus piperatus (Bull.) Bataille Chapip EcM 1 yes Cortinarius sp.01 Cor_01 EcM 1 yes Cortinarius sp.02 Cor_02 EcM 1 yes Cortinarius sp.03 Cor_03 EcM 1 yes Cortinarius sp.04 Cor_04 EcM 1 yes Cortinarius sp.05 Cor_05 EcM 1 yes Cortinarius sp.06 Cor_06 EcM 1 yes 57

Cortinarius sp.09 Cor_09 EcM 1 yes Cortinarius sp.10 Cor_10 EcM 1 yes Cortinarius sp.11 Cor_11 EcM 1 yes Cortinarius sp.12 Cor_12 EcM 1 yes Cortinarius sp.13 Cor_13 EcM 1 yes Cortinarius sp.16 Cor_16 EcM 1 yes Cortinarius sp.17 Cor_17 EcM 1 yes Cortinarius sp.18 Cor_18 EcM 1 yes Cortinarius sp.19 Cor_19 EcM 1 yes Cortinarius sp.20 Cor_20 EcM 1 yes Cortinarius sp.21 Cor_21 EcM 1 yes Cortinarius albocyaneus Fr. Coralc EcM 1 yes Cortinarius anomalus (Fr.) Fr. Corano EcM 1 yes Cortinarius anserinus (Velen.) Rob. Henry Corans EcM 1 yes Cortinarius barbatus (Batsch) Melot Corbar EcM 1 yes Cortinarius bataillei J. Favre Corbat EcM 1 yes Cortinarius camphoratus (Fr) Fr. Corcam EcM 1 yes Cortinarius caperatus (Pers.) Fr. Corcap EcM 1 yes Cortinarius cinnamomeus (L.) Gray Corcin EcM 1 yes Cortinarius citrinus (J.E. Lange) P.D. Orton Corcit EcM 1 yes Cortinarius comptulus M.M. Moser Corcom EcM 1 yes Cortinarius croceocaeruleus (Pers.) Fr. Corcrc EcM 1 yes Cortinarius delibutus Fr. Cordel EcM 1 yes Cortinarius depressus Fr. Cordep EcM 1 yes Cortinarius diasemospermus var. H. Lindstr. Cordil EcM 1 yes leptospermus Cortinarius flexipes var. inolens H. Lindstr. Corfli EcM 1 yes Cortinarius fulvescens s.l. Fr. Corful EcM 1 yes Cortinarius herpeticus Fr. Corher EcM 1 yes Cortinarius lebretonii Quél. Corleb EcM 1 yes Cortinarius obtusus (Fr.) Fr. Corobt EcM 1 yes Cortinarius raphanoides (Pers.) Fr. Corrap EcM 1 yes Cortinarius subbalaustinus Rob. Henry Corsbb EcM 1 yes Cortinarius subpurpurascens (Batsch) Fr. Corsbu EcM 1 yes Cortinarius scaurotraganoides Rob. Henry ex Rob. Corsca EcM 1 yes Henry Cortinarius semisanguineus (Fr.) Gillet Corsem EcM 1 yes Cortinarius turgidus Fr. Cortur EcM 1 yes Cortinarius uraceonemoralis Niskanen, Liimat., Dima, Corura EcM 1 yes Kytöv., Bojantchev & H. Lindstr. Cortinarius urbicus (Fr.) Fr. Corurb EcM 1 yes Cortinarius valgus Fr. Corval EcM 1 yes Cortinarius variecolor (Pers.) Fr. Corvar EcM 1 yes Cortinarius vulpinus (Velen.) Rob. Henry Corvul EcM 1 yes Cortinarius xanthocephalus P.D. Orton Corxcp EcM 1 yes Cortinarius xanthophyllus (Cooke) Rob. Henry Corxph EcM 1 yes Gomphidius roseus (Fr.) Fr. Gomros EcM 1 yes Hebeloma candidipes Bruchet Hebcan EcM 1 yes Hebeloma sacchariolens Quél. Hebsac EcM 1 yes Hydnum sp. Hyd_sp EcM 1 yes Hygrophorus agathosmus (Fr.) Fr. Hygaga EcM 1 yes Hygrophorus lindtneri M.M. Moser Hyglin EcM 1 yes Hygrophorus penarioides Jacobsson & E. Larss. Hygpen EcM 1 yes Hygrophorus unicolor Gröger Hyguni EcM 1 yes Inocybe amblyospora cf. Kühner Ino_am EcM 1 yes Inocybe auricoma cf. (Batsch) J.E. Lange Ino_au EcM 1 yes 58

Inocybe castanea Peck Inocas EcM 1 yes Inocybe flocculosa Sacc. Inoflo EcM 1 yes Inocybe grammata Quél. & Le Bret. Inogra EcM 1 yes Inocybe leiocephala D.E. Stuntz Inolei EcM 1 yes Inocybe microspora J.E. Lange Inomic EcM 1 yes Inocybe putilla Bres. Inoput EcM 1 yes Inocybe rimosa (Bull.) P. Kumm. Inorim EcM 1 yes Inocybe splendens R. Heim Inospl EcM 1 yes Lactarius bertillonii (Neuhoff ex Z. Schaef.) Lacber EcM 1 yes Bon Laccaria bicolor (Maire) P.D. Orton Lacbic EcM 1 yes Lactarius deterrimus Gröger Lacdet EcM 1 yes Lactarius glyciosmus (Fr.) Fr. Lacgly EcM 1 yes Lactarius pallidus Pers. Lacpal EcM 1 yes Lactarius vietus (Fr.) Fr. Lacvie EcM 1 yes Leccinum cyaneobasileucum Lannoy & Estad?s Leccya EcM 1 yes Leucocortinarius bulbiger (Alb. & Schwein.) Singer Leubul EcM 1 yes Phellodon melaleucus (Sw. ex Fr.) P. Karst. Phemel EcM 1 yes Ramaria fagetorum cf. Maas Geest. ex Schild Ram_fa EcM 1 yes Ramaria formosa (Pers.) Quél. Ramfor EcM 1 yes Rhizopogon roseolus (Corda) Th. Fr. Rhiros EcM 1 yes Russula aeruginea Lindblad Rusaer EcM 1 yes Russula amarissima Romagn. & E.-J. Gilbert Rusama EcM 1 yes Russula clavipes Velen. Ruscla EcM 1 yes Russula cremeoavellanea Singer Ruscre EcM 1 yes Russula farinipes Romell Rusfar EcM 1 yes Russula grisea Fr. Rusgri EcM 1 yes Russula lutensis Romagn. & Le Gal Ruslut EcM 1 yes Russula minutula Velen. Rusmin EcM 1 yes Russula nitida (Pers.) Fr. Rusnit EcM 1 yes Russula pseudointegra Arnould & Goris Ruspse EcM 1 yes Russula queletii Fr. Rusque EcM 1 yes Russula rhodella E.-J. Gilbert Rusrho EcM 1 yes Russula solaris Ferd. & Winge Russol EcM 1 yes Russula tinctipes J. Blum ex Bon Rustin EcM 1 yes Russula torulosa Bres. Rustor EcM 1 yes Russula virescens (Schaeff.) Fr. Rusvir EcM 1 yes Scleroderma cepa Pers. Sclcep EcM 1 yes Strobilomyces strobilaceus (Scop.) Berk. Strstr EcM 1 yes Suillus luteus (L.) Roussel Suilut EcM 1 yes Thelephora terrestris Ehrh. Theter EcM 1 yes Tricholoma batschii Gulden Tribat EcM 1 yes Tricholoma scalpturatum (Fr.) Quél. Trisca EcM 1 yes Tricholoma stiparophyllum (N. Lund) P. Karst. Tristi EcM 1 yes Xerocomus chrysonema A.E. Hills & A.F.S. Xerchr EcM 1 yes Taylor Xerocomus parasiticus (Bull.) Quél. Xerpar EcM 1 yes Xerocomus ripariellus Redeuilh Xerrip EcM 1 yes Auriscalpium vulgare Gray Aurvul t. sapr. 25 yes, plotted Lycoperdon perlatum Pers. Lycper t. sapr. 24 yes, plotted Mycena pura (Pers.) P. Kumm. Mycpur t. sapr. 22 yes, plotted Gymnopus peronatus (Bolton) Antonín, Halling Gymper t. sapr. 21 yes, plotted & Noordel. Leotia lubrica (Scop.) Pers. Leolub t. sapr. 20 yes, plotted Mycena sanguinolenta (Alb. & Schwein.) P. Mycsan t. sapr. 18 yes, plotted Kumm. Gymnopus aquosus (Bull.) Antonín & Gymaqu t. sapr. 17 yes, plotted 59

Noordel. Rhodocollybia butyracea (Bull.) Lennox Rhobut t. sapr. 16 yes, plotted Clitocybe nebularis (Batsch) P. Kumm. Clineb t. sapr. 15 yes, plotted Baeospora myosura (Fr.) Singer Baemyo t. sapr. 14 yes, plotted Lepista nuda (Bull.) Cooke Lepnud t. sapr. 14 yes, plotted Lycoperdon molle Pers. Lycmol t. sapr. 14 yes, plotted Mycena galopus var. galopus (Pers.) P. Kumm. Mycglp t. sapr. 13 yes Lepiota clypeolaria (Bull.) P. Kumm. Lepcly t. sapr. 12 yes Mycena aurantiomarginata (Fr.) Quél. Mycaur t. sapr. 12 yes Mycena flavescens Velen. Mycflv t. sapr. 12 yes Mycena rosea Gramberg Mycrsa t. sapr. 11 yes Clitocybe ditopa (Fr.) Gillet Clidit t. sapr. 10 yes Gymnopus erythropus (Pers.) Antonín, Halling Gymery t. sapr. 10 yes & Noordel. Mycena zephirus (Fr.) P. Kumm. Myczep t. sapr. 10 yes Lepista flaccida (Sowerby) Pat. Lepfla t. sapr. 9 yes Clitocybe candicans (Pers.) P. Kumm. Clican t. sapr. 8 yes Clitocybe phyllophila (Pers.) P. Kumm. Cliphy t. sapr. 8 yes Gymnopus androsaceus (L.) J.L. Mata & R.H. Gymand t. sapr. 8 yes Petersen Infundibulicybe gibba (Pers.) Harmaja Infgib t. sapr. 8 yes Lepiota castanea Quél. Lepcas t. sapr. 8 yes Marasmius bulliardii Quél. Marbul t. sapr. 8 yes Atheniella flavoalba (Fr.) Redhead, Moncalvo, Athfla t. sapr. 7 yes Vilgalys, Desjardin, B.A. Perry Lycoperdon nigrescens Pers. Lycnig t. sapr. 7 yes Lyophyllum platypum Kühner Lyopla t. sapr. 7 yes tenacellus (Pers.) Singer Strten t. sapr. 7 yes Collybia cirrata (Schumach.) Quél. Colcir t. sapr. 6 yes Conocybe tetrasporoides Hauskn. Contet t. sapr. 6 yes Hygrophoropsis aurantiaca (Wulfen) Maire Hygaur t. sapr. 6 yes Macrotyphula juncea (Alb. & Schwein.) Macjun t. sapr. 6 yes Berthier Macrolepiota procera (Scop.) Singer Macpro t. sapr. 6 yes Mycena amicta (Fr.) Quél. Mycami t. sapr. 6 yes Mycena galopus var. leucogala (Cooke) J.E. Lange Mycgll t. sapr. 6 yes Collybia tuberosa (Bull.) P. Kumm. Coltub t. sapr. 5 yes Entoloma juncinum (Kühner & Romagn.) Entjun t. sapr. 5 yes Noordel. Gymnopus confluens (Pers.) Antonín, Halling Gymcon t. sapr. 5 yes & Noordel. (Bull.) Murrill Gymdry t. sapr. 5 yes Lycoperdon excipuliforme (Scop.) Pers. Lycexc t. sapr. 5 yes Mycena rosella (Fr.) P. Kumm. Mycrla t. sapr. 5 yes Mycena stylobates (Pers.) P. Kumm. Mycsty t. sapr. 5 yes Roridomyces roridus (Scop.) Rexer Rorror t. sapr. 5 yes Tubaria minutalis Romagn. Tubmin t. sapr. 5 yes Clitocybe metachroa (Fr.) P. Kumm. Climet t. sapr. 4 yes Clitocybe phaeophthalma (Pers.) Kuyper Clipha t. sapr. 4 yes Collybia cookei (Bres.) J.D. Arnold Colcoo t. sapr. 4 yes Conocybe moseri Watling Conmos t. sapr. 4 yes Lyophyllum mephiticum (Fr.) Singer Lyomep t. sapr. 4 yes Mycena capillaris (Schumach.) P. Kumm. Myccap t. sapr. 4 yes Naucoria bohemica Velen. Nauboh t. sapr. 4 yes Ramaria flaccida (Fr.) Bourdot Ramfla t. sapr. 4 yes Agaricus essettei Bon Agaess t. sapr. 3 yes 60

Chlorophyllum olivieri (Barla) Vellinga Chloli t. sapr. 3 yes Clitocybe odora (Bull.) P. Kumm. Cliodo t. sapr. 3 yes Clitopilus prunulus (Scop.) P. Kumm. Clipru t. sapr. 3 yes Cystoderma amianthinum (Scop.) Fayod Cysami t. sapr. 3 yes Gymnopus quercophilus (Pouzar) Antonín & Gymque t. sapr. 3 yes Noordel. Helvella elastica Bull. Helela t. sapr. 3 yes Lepiota cristata (Bolton) P. Kumm. Lepcri t. sapr. 3 yes Lepiota ignivolvata Bousset & Joss. ex Joss. Lepign t. sapr. 3 yes Marasmius cohaerens (Pers.) Cooke & Quél. Marcoh t. sapr. 3 yes Mycena filopes (Bull.) P. Kumm. Mycfil t. sapr. 3 yes Mycena metata (Fr.) P. Kumm. Mycmet t. sapr. 3 yes Phallus impudicus L. Phaimp t. sapr. 3 yes Pholiotina brunnea (Watling) Singer Phobru t. sapr. 3 yes (Alb. & Schwein.) P. Riptri t. sapr. 3 yes Karst. Strobilurus stephanocystis (Kühner & Romagn. ex Strste t. sapr. 3 yes Hora) Singer Tubaria conspersa (Pers.) Fayod Tubcon t. sapr. 3 yes Tubaria furfuracea (Pers.) Gillet Tubfur t. sapr. 3 yes Agaricus semotus Fr. Agasem t. sapr. 2 yes Ampulloclitocybe clavipes (Pers.) Redhead, Lutzoni, Ampcla t. sapr. 2 yes Moncalvo & Vilgalys Anthina flammea (Jungh.) Fr. Antfla t. sapr. 2 yes Clitocybe fragrans (With.) P. Kumm. Clifra t. sapr. 2 yes Conocybe enderlei var. enderlei Hauskn. Conend t. sapr. 2 yes Cystolepiota seminuda (Lasch) Bon Cyssem t. sapr. 2 yes Entoloma hebes (Romagn.) Trimbach Entheb t. sapr. 2 yes Gymnopus ocior (Pers.) Antonín & Gymoci t. sapr. 2 yes Noordel. Helvella lacunosa Afzel. Hellac t. sapr. 2 yes Helvella macropus (Pers.) Gray Helmac t. sapr. 2 yes Lyophyllum leucophaeatum (P. Karst.) P. Karst. Lyoleu t. sapr. 2 yes Mycena abramsii cf. (Murrill) Murrill Myc_ab t. sapr. 2 yes Mycena fagetorum cf. (Fr.) Gillet Myc_fa t. sapr. 2 yes Mycena clavicularis (Fr.) Gillet Myccla t. sapr. 2 yes Mycena diosma Krieglst. & Schwöbel Mycdio t. sapr. 2 yes Mycena rubromarginata (Fr.) P. Kumm. Mycrub t. sapr. 2 yes Peziza saniosa Schrad. Pezsan t. sapr. 2 yes Rhodocybe gemina (Paulet) Kuyper & Rhogem t. sapr. 2 yes Noordel. Strobilurus esculentus (Wulfen) Singer Stresc t. sapr. 2 yes Agaricus sylvaticus Schaeff. Agasyl t. sapr. 1 yes Agrocybe vervacti (Fr.) Singer Agrver t. sapr. 1 yes Ciboria amentacea cf. (Balb.) Fuckel Cib_am t. sapr. 1 yes Clavariadelphus pistillaris (L.) Donk Clapis t. sapr. 1 yes Conocybe macrocephala cf. Kühner & Watling Con_ma t. sapr. 1 yes Conocybe ochrostriata var. Hauskn. Conovo t. sapr. 1 yes ochrostriata Coprinopsis jonesii (Peck) Redhead, Vilgalys Copjon t. sapr. 1 yes & Moncalvo Cystodermella cinnabarina (Alb. & Schwein.) Cyscin t. sapr. 1 yes Harmaja Entoloma conferendum var. (Velen.) Noordel. Entcon t. sapr. 1 yes pusillum Entoloma jahnii Wölfel & Winterh. – t. sapr. 1 no, omitted Lepiota boudieri Bres. Lepbou t. sapr. 1 yes 61

Lepista glaucocana (Bres.) Singer Lepgla t. sapr. 1 yes Lycoperdon lividum Pers. Lycliv t. sapr. 1 yes Lyophyllum baeospermum Romagn. Lyobae t. sapr. 1 yes Lyophyllum boudieri Kühner & Romagn. Lyobou t. sapr. 1 yes Lyophyllum rancidum (Fr.) Singer Lyoran t. sapr. 1 yes Macrocystidia cucumis (Pers.) Joss. Maccuc t. sapr. 1 yes Macrolepiota mastoidea (Fr.) Singer Macmas t. sapr. 1 yes Marasmius epiphyllus (Pers.) Fr. Marepi t. sapr. 1 yes Marasmius setosus (Sowerby) Noordel. Marset t. sapr. 1 yes Marasmius wynneae Berk. & Broome Marwyn t. sapr. 1 yes Mycena rebaudengi cf. Robich Myc_re t. sapr. 1 yes Mycena cinerella (P. Karst.) P. Karst. Myccin t. sapr. 1 yes Mycena pelianthina (Fr.) Quél. Mycpel t. sapr. 1 yes Mycena polyadelpha (Lasch) Kühner Mycpla t. sapr. 1 yes Mycena rhenana Maas Geest. & Winterh. Mycrhe t. sapr. 1 yes Mycena vulgaris (Pers.) P. Kumm. Mycvul t. sapr. 1 yes Mycetinis scorodonius (Fr.) A. Wilson & Mycsco t. sapr. 1 yes Desjardin Peziza arvernensis cf. Roze & Boud. Pez_ar t. sapr. 1 yes Peziza badia Pers. Pezbad t. sapr. 1 yes Peziza phyllogena Cooke Pezphy t. sapr. 1 yes Peziza succosa Berk. Pezsuc t. sapr. 1 yes Ramaria eumorpha (P. Karst.) Corner Rameum t. sapr. 1 yes Stropharia cyanea (Bull.) Tuom. Strcya t. sapr. 1 yes Tarzetta cupularis (L.) Svrček Tarcup t. sapr. 1 yes Stereum hirsutum (Willd.) Pers. Stehir wood-inh. 35 yes, plotted Exidia nigricans (With.) P. Roberts Exinig wood-inh. 31 yes, plotted Schizopora paradoxa s.l. (Schrad.) Donk Schpar wood-inh. 31 yes, plotted Schizopora flavipora (Berk. & M.A. Curtis ex Schfla wood-inh. 29 yes, plotted Cooke) Ryvarden Stereum ochraceoflavum (Schwein.) Sacc. Steocf wood-inh. 29 yes, plotted Steccherinum ochraceum (Pers.) Gray Steoch wood-inh. 26 yes, plotted Mycena vitilis (Fr.) Quél. Mycvit wood-inh. 25 yes, plotted Xylaria hypoxylon (L.) Grev. Xylhyp wood-inh. 25 yes, plotted Hymenochaete rubiginosa (Dicks.) Lév. Hymrub wood-inh. 24 yes, plotted Antrodiella fragrans (A. David & Tortič) A. Antfra wood-inh. 21 yes, plotted David & Tortič Skeletocutis nivea (Jungh.) Jean Keller Skeniv wood-inh. 21 yes, plotted Hymenopellis radicata (Relhan) R.H. Petersen Hymrad wood-inh. 19 yes, plotted Mycetinis alliaceus (Jacq.) Earle ex A.W. Mycall wood-inh. 19 yes, plotted Wilson & Desjardin Biscogniauxia nummularia (Bull.) Kuntze Bisnum wood-inh. 18 yes, plotted Mycena polygramma (Bull.) Gray Mycpgr wood-inh. 18 yes, plotted Stereum subtomentosum Pouzar Stesub wood-inh. 18 yes, plotted Crepidotus variabilis (Pers.) P. Kumm. Crevar wood-inh. 17 yes, plotted Galerina marginata (Batsch) Kühner Galmar wood-inh. 16 yes, plotted Hypocrea citrina (Pers.) Fr. Hypcit wood-inh. 16 yes, plotted Lycoperdon pyriforme Willd. Lycpyr wood-inh. 16 yes, plotted Mycena epipterygia (Scop.) Gray Mycepi wood-inh. 16 yes, plotted Ramaria stricta (Pers.) Quél. Ramstr wood-inh. 16 yes, plotted Stereum sanguinolentum (Alb. & Schwein.) Fr. Stesan wood-inh. 16 yes, plotted Trametes versicolor (L.) Pilát Traver wood-inh. 16 yes, plotted Hypholoma fasciculare (Huds.) P. Kumm. Hypfas wood-inh. 15 yes, plotted Postia subcaesia (A. David) Jülich Possub wood-inh. 15 yes, plotted Schizophyllum commune Fr. Schcom wood-inh. 15 yes, plotted Xylaria carpophila (Pers.) Fr. Xylcar wood-inh. 15 yes, plotted Calocera furcata (Fr.) Fr. Calfur wood-inh. 14 yes, plotted 62

Hypoxylon fragiforme (Pers.) J. Kickx f. Hypfra wood-inh. 14 yes, plotted Mycena galericulata (Scop.) Gray Mycglu wood-inh. 14 yes, plotted Polyporus varius (Pers.) Fr. Polvar wood-inh. 14 yes, plotted Cyathus striatus (Huds.) Willd. Cyastr wood-inh. 13 yes Pluteus cervinus (Schaeff.) P. Kumm. Plucer wood-inh. 13 yes Postia stiptica (Pers.) Jülich Possti wood-inh. 13 yes Trametes hirsuta (Wulfen) Lloyd Trahir wood-inh. 13 yes Clitocybula platyphylla (Pers.) Malençon & Clipla wood-inh. 12 yes Bertault Diatrype stigma (Hoffm.) Fr. Diasti wood-inh. 12 yes Fuscoporia contigua (Pers.) G. Cunn. Fuscon wood-inh. 12 yes Panellus stipticus (Bull.) P. Karst. Pansti wood-inh. 12 yes Aleurodiscus disciformis (DC.) Pat. Aledis wood-inh. 11 yes Armillaria lutea Gillet Armlut wood-inh. 11 yes Exidia glandulosa (Bull.) Fr. Exigla wood-inh. 11 yes Junghuhnia nitida (Pers.) Ryvarden Junnit wood-inh. 11 yes Mycena maculata P. Karst. Mycmac wood-inh. 11 yes rufa (Pers.) M.P. Christ. Phlruf wood-inh. 11 yes Psathyrella pygmaea (Bull.) Singer Psapyg wood-inh. 10 yes Heterobasidion annosum (Fr.) Bref. Hetann wood-inh. 9 yes Laxitextum bicolor (Pers.) Lentz Laxbic wood-inh. 9 yes Mycena haematopus (Pers.) P. Kumm. Mychae wood-inh. 9 yes Plicaturopsis crispa (Pers.) D.A. Reid Plicri wood-inh. 9 yes Pseudohydnum gelatinosum (Scop.) P. Karst. Psegel wood-inh. 9 yes Antrodia albida (Fr.) Donk Antalb wood-inh. 8 yes Antrodia malicola (Berk. & M.A. Curtis) Antmal wood-inh. 8 yes Donk Auricularia auricula-judae (Bull.) Quél. Auraur wood-inh. 8 yes Crepidotus cesatii (Rabenh.) Sacc. Creces wood-inh. 8 yes Galerina pruinatipes A.H. Sm. Galpru wood-inh. 8 yes Gymnopilus penetrans (Fr.) Murrill Gympen wood-inh. 8 yes Hypholoma lateritium (Schaeff.) P. Kumm. Hyplat wood-inh. 8 yes lenta (Pers.) Singer Pholen wood-inh. 8 yes Polyporus alveolaris (DC.) Bondartsev & Polalv wood-inh. 8 yes Singer Simocybe centunculus (Fr.) Singer Simcen wood-inh. 8 yes Steccherinum fimbriatum (Pers.) J. Erikss. Stefim wood-inh. 8 yes (Pers.) Grev. Xylpol wood-inh. 8 yes Antrodiella faginea Vampola & Pouzar Antfag wood-inh. 7 yes Byssomerulius corium (Pers.) Parmasto Byscor wood-inh. 7 yes Calocera viscosa (Pers.) Fr. Calvis wood-inh. 7 yes Mycena arcangeliana Bres. Mycarc wood-inh. 7 yes Psathyrella piluliformis (Bull.) P.D. Orton Psapil wood-inh. 7 yes Bjerkandera adusta (Willd.) P. Karst. Bjeadu wood-inh. 6 yes Ceriporiopsis mucida (Pers.) Gilb. & Ryvarden Cermuc wood-inh. 6 yes Daedaleopsis confragosa (Bolton) J. Schröt. Daecon wood-inh. 6 yes Diatrype disciformis (Hoffm.) Fr. Diadis wood-inh. 6 yes Marasmiellus ramealis (Bull.) Singer Marram wood-inh. 6 yes Mycena inclinata (Fr.) Quél. Mycinc wood-inh. 6 yes Phellinus viticola (Schwein.) Donk Phevit wood-inh. 6 yes Polyporus ciliatus Fr. Polcil wood-inh. 6 yes Trichaptum abietinum (Dicks.) Ryvarden Triabi wood-inh. 6 yes Xylaria longipes Nitschke Xyllon wood-inh. 6 yes Antrodiella pallescens (Pilát) Niemelä & Antpal wood-inh. 5 yes Miettinen Calocera cornea (Batsch) Fr. Calcor wood-inh. 5 yes Hapalopilus nidulans (Fr.) P. Karst. Hapnid wood-inh. 5 yes 63

Mucidula mucida (Schrad.) Pat. Mucmuc wood-inh. 5 yes Phlebia radiata Fr. Phlrad wood-inh. 5 yes Piptoporus betulinus (Bull.) P. Karst. Pipbet wood-inh. 5 yes Pluteus semibulbosus (Lasch) Quél. Plusem wood-inh. 5 yes Postia caesia (Schrad.) P. Karst. Poscae wood-inh. 5 yes (Batsch) Gray Resapp wood-inh. 5 yes Trichaptum biforme (Fr.) Ryvarden Tribif wood-inh. 5 yes Annulohypoxylon multiforme (Fr.) Y.-M. Ju, J.D. Annmul wood-inh. 4 yes Rogers & H.-M. Hsieh Ascocoryne cylichnium (Tul.) Korf Asccyl wood-inh. 4 yes Ascocoryne sarcoides (Jacq.) J.W. Groves & Ascsar wood-inh. 4 yes D.E. Wilson Chlorociboria aeruginascens (Nyl.) Kanouse ex C.S. Chlaer wood-inh. 4 yes Ramamurthi, Korf & L.R. Batra Coprinellus micaceus (Bull.) Vilgalys, Hopple Copmic wood-inh. 4 yes & Jacq. Johnson (Schaeff.) Staude Cremol wood-inh. 4 yes Dacrymyces capitatus Schwein. Daccap wood-inh. 4 yes Diatrypella favacea (Fr.) Ces. & De Not. Diafav wood-inh. 4 yes Gymnopilus sapineus (Fr.) Murrill Gymsap wood-inh. 4 yes Hypoxylon rubiginosum (Pers.) Fr. Hyprub wood-inh. 4 yes Hypocrea rufa (Pers.) Fr. Hypruf wood-inh. 4 yes Lentinellus ursinus (Fr.) Kühner Lenurs wood-inh. 4 yes Mensularia nodulosa (Fr.) T. Wagner & M. Mennod wood-inh. 4 yes Fisch. Nemania serpens (Pers.) Gray Nemser wood-inh. 4 yes Peziza micropus Pers. Pezmic wood-inh. 4 yes Polyporus brumalis (Pers.) Fr. Polbru wood-inh. 4 yes Porotheleum fimbriatum (Pers.) Fr. Porfim wood-inh. 4 yes Postia tephroleuca (Fr.) Jülich Postep wood-inh. 4 yes Rutstroemia firma (Pers.) P. Karst. Rutfir wood-inh. 4 yes Sparassis crispa (Wulfen) Fr. Spacri wood-inh. 4 yes Steccherinum bourdotii Saliba & A. David Stebou wood-inh. 4 yes Stereum gausapatum (Fr.) Fr. Stegau wood-inh. 4 yes Tremella foliacea Pers. Trefol wood-inh. 4 yes Tremella mesenterica Retz. Tremes wood-inh. 4 yes Agrocybe praecox (Pers.) Fayod Agrpra wood-inh. 3 yes Annulohypoxylon cohaerens (Pers.) Y.-M. Ju, J.D. Anncoh wood-inh. 3 yes Rogers & H.-M. Hsieh Antrodiella romellii (Donk) Niemelä Antrom wood-inh. 3 yes Bulgaria inquinans (Pers.) Fr. Bulinq wood-inh. 3 yes Ceriporia purpurea (Fr.) Donk Cerpur wood-inh. 3 yes Dacrymyces stillatus Nees Dacsti wood-inh. 3 yes Daedalea quercina (L.) Pers. Daeque wood-inh. 3 yes Fuscoporia ferruginosa (Schrad.) Murrill Fusfrr wood-inh. 3 yes Galerina sideroides (Bull.) Kühner Galsid wood-inh. 3 yes Hypoxylon fuscum (Pers.) Fr. Hypfus wood-inh. 3 yes Hypocrea gelatinosa (Tode) Fr. Hypgel wood-inh. 3 yes Lenzites betulina (L.) Fr. Lenbet wood-inh. 3 yes Mycena stipata Maas Geest. & Schwöbel Mycsti wood-inh. 3 yes Oxyporus latemarginatus (Durieu & Mont.) Donk Oxylat wood-inh. 3 yes Phlebia livida (Pers.) Bres. Phlliv wood-inh. 3 yes Phlebia tremellosa (Schrad.) Nakasone & Phltre wood-inh. 3 yes Burds. Polyporus arcularius (Batsch) Fr. Polarc wood-inh. 3 yes Polyporus tuberaster (Jacq. ex Pers.) Fr. Poltub wood-inh. 3 yes 64

Postia fragilis (Fr.) Jülich Posfra wood-inh. 3 yes Postia simanii (Pilát ex Pilát) Jülich Possim wood-inh. 3 yes Ramaria apiculata (Fr.) Donk Ramapi wood-inh. 3 yes Skeletocutis amorpha (Fr.) Kotl. & Pouzar Skeamo wood-inh. 3 yes Tricholomopsis rutilans (Schaeff.) Singer Trirut wood-inh. 3 yes Antrodiella serpula (P. Karst.) Spirin & Antser wood-inh. 2 yes Niemelä Callistosporium luteo-olivaceum (Berk. & M.A. Curtis) Callut wood-inh. 2 yes Singer Coriolopsis trogii (Berk.) Domański Cortro wood-inh. 2 yes Crepidotus epibryus (Fr.) Quél. Creepi wood-inh. 2 yes Crepidotus luteolus Sacc. Crelut wood-inh. 2 yes Dacrymyces chrysospermus Berk. & M.A. Curtis Dacchr wood-inh. 2 yes Daedaleopsis tricolor (Bull.) Bond. & Sing. Daetri wood-inh. 2 yes Dentipellis fragilis (Pers.) Donk Denfra wood-inh. 2 yes Diatrypella quercina (Pers.) Cooke Diaque wood-inh. 2 yes Fomes fomentarius (L.) Fr. Fomfom wood-inh. 2 yes Ganoderma applanatum (Pers.) Pat. Ganapp wood-inh. 2 yes Gymnopus fusipes (Bull.) Gray Gymfus wood-inh. 2 yes Hypholoma capnoides (Fr.) P. Kumm. Hypcap wood-inh. 2 yes Irpex lacteus (Fr.) Fr. Irplac wood-inh. 2 yes Kretzschmaria deusta (Hoffm.) P.M.D. Martin Kredeu wood-inh. 2 yes Mycena crocata (Schrad.) P. Kumm. Myccro wood-inh. 2 yes Mycena hiemalis (Osbeck) Quél. Mychie wood-inh. 2 yes Mycena silvae-nigrae Maas Geest. & Schwöbel Mycsil wood-inh. 2 yes Phellinidium ferrugineofuscum (P. Karst.) Fiasson & Phefer wood-inh. 2 yes Niemelä Phloeomana speirea (Fr.) Redhead Phlspe wood-inh. 2 yes Physisporinus vitreus (Pers.) P. Karst. Phyvit wood-inh. 2 yes Pluteus leoninus (Schaeff.) P. Kumm. Pluleo wood-inh. 2 yes Pluteus podospileus Sacc. & Cub. Plupod wood-inh. 2 yes Pluteus thomsonii (Berk. & Broome) Dennis Plutho wood-inh. 2 yes Porostereum spadiceum (Pers.) Hjortstam & Porspa wood-inh. 2 yes Ryvarden Postia ptychogaster (F. Ludw.) Westerh. Pospty wood-inh. 2 yes Psathyrella cernua (Vahl) G. Hirsch Psacer wood-inh. 2 yes Psathyrella gossypina (Bull.) A. Pearson & Psagos wood-inh. 2 yes Dennis Pseudomerulius aureus (Fr.) Jülich Pseaur wood-inh. 2 yes Skeletocutis lenis (P. Karst.) Niemelä Skelen wood-inh. 2 yes Steccherinum cremeoalbum Hjortstam Stecre wood-inh. 2 yes Tapinella atrotomentosa (Batsch) Šutara Tapatr wood-inh. 2 yes Trametes gibbosa (Pers.) Fr. Tragib wood-inh. 2 yes Trechispora mollusca (Pers.) Liberta Tremol wood-inh. 2 yes Trichaptum fuscoviolaceum (Ehrenb.) Ryvarden Trifus wood-inh. 2 yes Agrocybe firma (Peck) Singer Agrfir wood-inh. 1 yes Antrodia vaillantii (DC.) Ryvarden Antvai wood-inh. 1 yes Armillaria mellea (Vahl) P. Kumm. Armmel wood-inh. 1 yes Armillaria ostoyae (Romagn.) Herink Armost wood-inh. 1 yes Artomyces pyxidatus (Pers.) Jülich Artpyx wood-inh. 1 yes Ascotremella faginea (Peck) Seaver Ascfag wood-inh. 1 yes Bjerkandera fumosa (Pers.) P. Karst. Bjefum wood-inh. 1 yes Bolbitius reticulatus (Pers.) Ricken Blbret wood-inh. 1 yes Bolbitius pluteoides M.M. Moser Bolplu wood-inh. 1 yes Cantharellula umbonata (J.F. Gmel.) Singer Canumb wood-inh. 1 yes Ceriporiopsis gilvescens (Bres.) Dom. Cergil wood-inh. 1 yes Cerrena unicolor (Bull.) Murrill Ceruni wood-inh. 1 yes 65

Crepidotus applanatus (Pers.) P. Kumm. Creapp wood-inh. 1 yes Crepidotus calolepis (Fr.) P. Karst. Crecal wood-inh. 1 yes Crepidotus versutus (Peck) Sacc. Crever wood-inh. 1 yes Cylindrobasidium laeve (Pers.) Chamuris Cyllae wood-inh. 1 yes Dacrymyces lacrymalis (Pers.) Sommerf. Daclac wood-inh. 1 yes Daldinia concentrica (Bolton) Ces. & De Not. Dalcon wood-inh. 1 yes Datronia mollis (Sommerf.) Donk Datmol wood-inh. 1 yes Deconica inquilina (Fr.) Romagn. Decinq wood-inh. 1 yes Dichomitus campestris (Quél.) Dom. & Orlicz Diccam wood-inh. 1 yes Flammulaster carpophilus (Fr.) Earle Flacar wood-inh. 1 yes Flammulaster limulatus var. lituus Vellinga Flalim wood-inh. 1 yes Fomitiporia punctata (P. Karst.) Murrill Fompun wood-inh. 1 yes Fomitiporia robusta (P. Karst.) Fiasson & Fomrob wood-inh. 1 yes Niemelä Galerina camerina cf. (Fr.) Kühner Gal_ca wood-inh. 1 yes Galerina pallida cf. (Pilát) E. Horak & M.M. Gal_pa wood-inh. 1 yes Moser Galerina triscopa (Fr.) Kühner Galtri wood-inh. 1 yes Gloeoporus dichrous (Fr.) Bres. Glodic wood-inh. 1 yes Guepiniopsis buccina (Pers.) L.L. Kenn. Guebuc wood-inh. 1 yes Gyromitra infula (Schaeff.) Quél. Gyrinf wood-inh. 1 yes Hydropus subalpinus (Höhn.) Singer Hydsub wood-inh. 1 yes Hypoxylon ferrugineum cf. G.H. Otth Hyp_fe wood-inh. 1 yes Hypoxylon howeanum Peck Hyphow wood-inh. 1 yes Hypocrea sulphurea (Schwein.) Sacc. Hypsul wood-inh. 1 yes Inonotus nidus-pici Pilát Inonid wood-inh. 1 yes Lentinellus cochleatus (Pers.) P. Karst. Lencoc wood-inh. 1 yes Lentinellus flabelliformis (Bolton) S. Ito Lenfla wood-inh. 1 yes Mycoacia aurea (Fr.) J. Erikss. & Mycaua wood-inh. 1 yes Ryvarden Mycena erubescens Höhn. Myceru wood-inh. 1 yes Mycoacia uda (Fr.) Donk Mycuda wood-inh. 1 yes Mycena viridimarginata P. Karst. Mycvir wood-inh. 1 yes Nemania atropurpurea (Fr.) Pouzar Nematr wood-inh. 1 yes Oxyporus obducens cf. (Pers.) Donk Oxy_ob wood-inh. 1 yes Oxyporus populinus (Schumach.) Donk Oxypop wood-inh. 1 yes Phaeomarasmius erinaceus (Fr.) Scherff. ex Romagn. Phaeri wood-inh. 1 yes Phaeolus schweinitzii (Fr.) Pat. Phasch wood-inh. 1 yes Phellinus pomaceus (Pers.) Maire Phepom wood-inh. 1 yes Phellinus tremulae (Bondartsev) Bondartsev Phetre wood-inh. 1 yes & P.N. Borisov Phlebiella vaga (Fr.) P. Karst. Phlvag wood-inh. 1 yes Pholiota flammans (Batsch) P. Kumm. Phofla wood-inh. 1 yes Pholiota gummosa (Lasch) Singer Phogum wood-inh. 1 yes Pholiota jahnii Tjall.-Beuk. & Bas Phojah wood-inh. 1 yes Pholiota spumosa (Fr.) Singer Phospu wood-inh. 1 yes pulmonarius (Fr.) Quél. Plepul wood-inh. 1 yes Pluteus exiguus (Pat.) Sacc. Pluexi wood-inh. 1 yes Pluteus nanus (Pers.) P. Kumm. Plunan wood-inh. 1 yes Pluteus pellitus (Pers.) P. Kumm. Plupel wood-inh. 1 yes Pluteus romellii (Britzelm.) Sacc. Plurom wood-inh. 1 yes Pluteus salicinus (Pers.) P. Kumm. Plusal wood-inh. 1 yes Pluteus satur Kühner & Romagn. Plusat wood-inh. 1 yes Psathyrella olympiana cf. A.H. Sm. Psa_ol wood-inh. 1 yes Resupinatus trichotis (Pers.) Singer Restri wood-inh. 1 yes Rigidoporus sanguinolentus (Alb. & Schwein.) Donk Rigsan wood-inh. 1 yes Skeletocutis alutacea cf. (J. Lowe) Jean Keller Ske_al wood-inh. 1 yes 66

Skeletocutis carneogrisea A. David Skecar wood-inh. 1 yes Stereum rugosum Pers. Sterug wood-inh. 1 yes Tapinella panuoides (Fr.) E.-J. Gilbert Tappan wood-inh. 1 yes Trametopsis cervina (Schwein.) Tomšovský Tracer wood-inh. 1 yes Trametes ochracea (Pers.) Gilb. & Ryvarden Traoch wood-inh. 1 yes Trametes suaveolens (L.) Fr. Trasua wood-inh. 1 yes Tyromyces chioneus (Fr.) P. Karst. Tyrchi wood-inh. 1 yes Volvariella caesiotincta P.D. Orton Volcae wood-inh. 1 yes Xerula pudens (Pers.) Singer Xerpud wood-inh. 1 yes Rickenella fibula (Bull.) Raithelh. – bryophilous 8 no, skipped Rickenella swartzii (Fr.) Kuyper – bryophilous 4 no, skipped Cordyceps larvicola Quél. – entomopath. 1 no, skipped Marasmius rotula (Scop.) Fr. – lign./t. sapr. 6 no, skipped Marasmius torquescens Quél. – lign./t. sapr. 5 no, skipped Mycena leptocephala (Pers.) Gillet – lign./t. sapr. 4 no, skipped Psathyrella lutensis (Romagn.) Bon – lign./t. sapr. 3 no, skipped Pholiota scamba (Fr.) M.M. Moser – lign./t. sapr. 1 no, skipped Psathyrella cortinarioides P.D. Orton – lign./t. sapr. 1 no, skipped Psathyrella fagetophila Örstadius & Enderle – lign./t. sapr. 1 no, skipped Psathyrella microrrhiza (Lasch) Konrad & Maubl. – lign./t. sapr. 1 no, skipped Psathyrella prona (Fr.) Gillet – lign./t. sapr. 1 no, skipped Psathyrella spadiceogrisea (Schaeff.) Maire – lign./t. sapr. 1 no, skipped Asterophora lycoperdoides (Bull.) Ditmar – mycotrophic 7 no, skipped Tremella encephala Pers. – mycotrophic 6 no, skipped Elaphocordyceps ophioglossoides (Ehrh.) G.H. Sung, J.M. – mycotrophic 3 no, skipped Sung & Spatafora Tremella globispora D.A. Reid – mycotrophic 1 no, skipped Entoloma rhodopolium (Fr.) P. Kumm. – t. sapr./myc. 15 no, skipped Otidea onotica (Pers.) Fuckel – t. sapr./myc. 11 no, skipped Otidea alutacea (Pers.) Massee – t. sapr./myc. 5 no, skipped Entoloma politum (Pers.) Donk – t. sapr./myc. 4 no, skipped Otidea bufonia (Pers.) Boud. – t. sapr./myc. 4 no, skipped Otidea fuckelii M. Carbone & Van – t. sapr./myc. 1 no, skipped Vooren Otidea grandis (Pers.) Arnould – t. sapr./myc. 1 no, skipped Otidea propinquata cf. (P. Karst.) Harmaja – t. sapr./myc. 1 no, skipped

67

Supplementary Fig 1. NMDS results of wood-inhabiting macrofungi representing the optimal positions of all the 245 collected species (red letters; see Supplementary Table 1 for abbreviations), the significantly (p < 0.05; based on 999 replications) fitted environmental variables (blue capitals), and a tri-plot of 35 (all) sampling units (black numbers). By seeking a stable NMDS result with a low final stress, a 3-D solution was chosen: in Fig (a) axis 1 of the ordination is plotted against axis 2; while in Fig (b) the first and the third axes of the same run are shown. As a determination of goodness of fit, a Shepard diagram is drawn in Fig (c) where ordination distances are plotted against the observed dissimilarities and 20 fits are shown as monotonic step lines representing each run after which the best NMDS solution was reached (non-metric fit: r2 = 0.977; linear fit: r2 = 0.748). Random starting configurations were used for finding the best solution. Fig (d) displays the same diagram, but with the best- fit monotonic regression of distances. The red line denotes hypothetical distances that would be in the perfect rank-order with the dissimilarities (scatter about this line defines the NMDS stress). Fig (e) was used to select the optimal dimensionality where the Kendall’s rank correlation coefficients (τ), indicating that how good the original distance matrix was recovered by the ordination distances, were plotted against the final NMDS stress values. Ten dimensions (black numbers) were tested in total. To the 3-D solution chosen, τ = 0.5114 was related. The 4-D final solution with a better τ and lower stress was avoided because of the decreased interpretability of the results. Distance method applied: Bray–Curtis; final stress: 15.282 using Kruskal’s stress formula 1 multiplied by 100 (Kruskal, 1964).

68

(a)

69

(b)

70

(c) (d) (e)

NMDS stress NMDS

Ordination distance Ordination

Observed dissimilarity Kendall’s τ

71

Supplementary Fig 2. NMDS results of terricolous saprotrophic macrofungi showing the optimal positions of 126 collected species (red letters; labels are coded as per Supplementary Table 1), the significantly (p < 0.05) fitted environmental variables (based on 999 replications, blue capitals), and a tri-plot of 31 sampling units (black numbers). Four sampling units (namely 111, 117, 119 and 120) with zero or very low species counts were excluded as these can have a disproportionate effect on the results. For finding a stable NMDS result with a low final stress, a 3-D solution was carried out: in Fig (a) the first two NMDS axes; while in Fig (b) the first and the third axis of the same run are plotted. In Fig (c), goodness of fit was mapped by the Shepard diagram of the NMDS result where ordination distances are plotted against the observed dissimilarities and 20 fits are shown as monotonic step lines representing each run after which the best NMDS solution was reached (non-metric fit: r2 = 0.979; linear fit: r2 = 0.82). Random starting configurations were used for finding the best solution. Fig (d) displays the same diagram, but with the best-fit monotonic regression of distances. The red line denotes hypothetical distances that would be in the perfect rank-order with the dissimilarities (scatter about this line defines the NMDS stress). A large step is demonstrated in the diagram because despite the fact that four sites had already been exluded, some sites (especially plot 137) did have a number of no shared or relatively rare species. To handle these tied dissimilarity values adequately, the function “step-across dissimilarities” were used in the package “vegan” for improving the NMDS results. Fig (e) was used to select the optimal dimensionality where the Kendall’s rank correlation coefficients (τ), indicating that how good the original distance matrix was recovered by the ordination distances, were plotted against the final NMDS stress values. Ten dimensions (black numbers) were tested in total. To the 3-D solution chosen, τ = 0.5864 was related. Distance method applied: Bray–Curtis; final stress: 14.331 using Kruskal’s stress formula 1 multiplied by 100 (Kruskal, 1964). (a)

72

73

(b)

(c) (d) (e)

NMDS stress NMDS

Ordination distance Ordination

74

Observed dissimilarity Kendall’s τ

75

Supplementary Fig 3. NMDS results of ectomycorrhizal macrofungi showing the optimal positions of all the 290 collected species (red letters; legend in Supplementary Table 1), the significantly (p < 0.05) fitted environmental variables (based on 999 replications and depicted by blue arrows), and a tri-plot of 35 (all) sampling units (black numbers). NMDS reached a stable solution with the lowest final stress by adding a third dimension. In Fig (a), NMDS axis 1 against axis 2, while in Fig (b) axis 1 against axis 3 are plotted. For determining goodness of fit, a Shepard diagram was displayed in Fig (c) where ordination distances are plotted against the observed dissimilarities with 20 fits as monotonic step lines representing each run after which the best NMDS solution was reached (non-metric fit: r2 = 0.985; linear fit: r2 = 0.846). Random starting configurations were used for finding the best solution. Fig (d) reports the same diagram, but with the best-fit monotonic regression of distances. The red line denotes hypothetical distances that would be in the perfect rank-order with the dissimilarities (scatter about this line defines the NMDS stress). Fig (e) was used to select the optimal dimensionality where the Kendall’s rank correlation coefficients (τ), indicating that how good the original distance matrix was recovered by the ordination distances, were plotted against the final NMDS stress values. Ten dimensions (black numbers) were tested in total. To the 3- D solution chosen, τ = 0.6165 was related. Distance method used: Bray–Curtis; final stress: 12.186 applying Kruskal’s stress formula 1 multiplied by 100 (Kruskal, 1964).

76

(a)

77

(b)

78

(c) (d) (e)

NMDS stress NMDS

Ordination distance Ordination

Observed dissimilarity Kendall’ s τ

79

Supplementary Fig 4. Redundancy analysis (RDA) of wood-inhabiting fungi. In Fig 2a of the printed paper, the same taxa are plotted in the NMDS diagram. Applying RDA, the (rare) species collected in less than four sampling units were omitted (see Supplementary Table 1). All of the more frequent taxa (black italics) were accepted for building RDA axes. Compared to the NMDS results, RDA highlighted very similar environmental factors (red letters) to be important for the species composition of wood-inhabiting fungi. RDA resulted eight variables with significant (p < 0.05) effects showing the relative volumes of dominant tree species as of the greatest importance. Scots pine and beech (including hornbeam) revealed a clear deciduous–coniferous gradient. The canonical axes explained 37.4% of the total variance. The majority of plotted species had positive scores along axis 1 preferring high proportions of deciduous trees. Fungi species with high scores and strong relations to beech (deciduous stands) in RDA were Antrodiella fragrans, Biscogniauxia nummularia, Mycetinis alliaceus, Postia subcaesia, Skeletocutis nivea, Xylaria carpophila and X. hypoxylon. Wood-inhabiting taxa in warmer, deciduous stands (dominated by oak species based on the scatter of sites) were Hymenochaete rubiginosa, Schizopora paradoxa s.l., Stereum ochraceoflavum and S. subtomentosum (sampling units are not shown). In relatively cool, pine-dominated stands Mycena epipterygia, Crepidotus variabilis and Stereum sanguinolentum were common. Dead wood related variables had no significant effects, and air temperature was not significant in the NMDS model. Wood-inhabiting fungi was the only functional group that was related significantly to a variable belonging to historical forest management practices.

Significance of all canonical axes: F = 2.279, p = 0.001. Explained variances of axes as percentages are shown. Percentage variance explained by the variables within the RDA model are listed.

Environmental variable Variance (%) F-value p-value Beech (relative volume of beech) 9.6 3.86 0.001 80

Scots pine (relative volume of Scots pine) 6.3 2.68 0.001 Temperature (mean daily air temperature difference) 5.0 2.19 0.001 Hornbeam (relative volume of hornbeam) 3.8 1.71 0.004 Non-dom. trees (relative volume of non-dominant trees) 3.4 1.56 0.022 Mean DBH (mean Diameter at Breast Height of trees) 3.3 1.57 0.017 pH of litter 3.0 1.46 0.034 Hist. meadows (historical proportion of meadows) 3.0 1.45 0.041

81

Supplementary Fig 5. Redundancy analysis (RDA) of terricolous saprotrophic fungi. In Fig 2b of the printed paper, the same taxa are plotted in the NMDS diagram. Applying RDA, the (rare) species collected in less than four sampling units were omitted (details in Supplementary Table 1). All the more frequent taxa were accepted for building RDA axes. All of the 35 sampling units were examined by RDA (sampling units are not shown), while NMDS was run by the omission of four sampling units with zero or very low counts of terricolous saprotrophic fungi. Broadly speaking, both methods revealed similar results: a definite litter pH gradient along the relative volume of Scots pine and the pH of litter (red letters). On the contrary, the effect of mean daily air temperature, however, was quite important in RDA, but it had no significant (p < 0.05) effect applying NMDS. In RDA, the canonical axes explained 31.6% of the total variance. Regarding the species, both methods highlighted Auriscalpium vulgare and Baeospora myosura to be common elements of pine- dominated stands. Using RDA, Lycoperdon perlatum and Gymnopus peronatus had relatively strong and positive relations to litter pH and the mass proportion of decayed litter, but the latter variable had no significant effect in the NMDS results (the position of Lycoperdon perlatum in the NMDS plot constructs an angle to litter pH very close to 90° showing no correlation with it). Supported by RDA only, stands with low air temperature and low soil N content were associated with the majority of frequent species: Clitocybe nebularis, Gymnopus aquosus, Leotia lubrica, Lepista nuda, Mycena pura, M. sanguinolenta, and Rhodocollybia butyracea (a). Moreover, the preponderance of all studied taxa preferred low air temperatures and low soil N contents (b).

(a) (b)

82

Significance of all canonical axes: F = 3.310, p = 0.001. Explained variances of axes as percentages are shown. Percentage variance explained by the variables within the RDA model are listed.

Environmental variable Variance (%) F-value p-value Scots pine (relative volume of Scots pine) 11.7 5.11 0.001 Temperature (mean daily air temperature difference) 9.1 4.37 0.001 pH of litter 3.6 1.84 0.008 Decayed litter (mass proportion of decayed litter) 3.6 1.77 0.018 Soil nitrogen (nitrogen content of soil) 3.5 1.86 0.006

83

Supplementary Fig 6. Redundancy analysis (RDA) of EcM fungi. In Fig 2c of the printed paper, the same taxa are plotted by using NMDS. Applying RDA, the (rare) species collected in less than four sampling units were omitted (details in Supplementary Table 1), but all of the more frequent taxa (black italics) were accepted for building RDA axes. The canonical axes explained 34.3% of the total variance, and six environmental factors were significant with the strongest effects of the relative volume of beech and the mean DBH of trees. The determination of RDA axis 1 was threefold: the relative volume of beech and the proportion of forests in the landscape correlated positively with it, while the mean of relative diffuse light correlated negatively with axis 1. The mean DBH of trees and the pH of litter had positive effects, whereas the diversity of landscape elements had a negative effect along axis 2. Axis 1 explained ca. three times more variation compared to axis 2; and most of the plotted species had strong (positive) correlations with it. The species that preferred closed beech stands with more neutral litter pH and high proportion of forests in the landscape were Inocybe petiginosa, Lactarius blennius, L. subdulcis, Pseudocraterellus undulatus, Tricholoma sulphureum and T. ustale. The stands with high mean DBH of trees were favoured by Clavulina cinerea, Humaria hemisphaerica and Laccaria laccata, while Russula fragilis and Lactarius chrysorrheus preferred stands with a low tree DBH and a high landscape diversity. Characteristic EcM taxa in open stands were Amanita rubescens, Lactarius quietus, Russula heterophylla, R. nigricans and R. undulata.

Significance of all canonical axes: F = 2.652, p = 0.001. Explained variances of axes as percentages are shown. Percentage variance explained by the variables within the RDA model are listed. 84

Environmental variable Variance (%) F-value p-value Beech (relative volume of beech) 8.8 3.32 0.001 Mean DBH (mean Diameter at Breast Height of trees) 6.6 2.65 0.003 Forests (proportion of forests in the landscape) 5.6 2.33 0.007 Light (mean relative diffuse light) 4.9 2.24 0.005 pH of litter 4.9 2.13 0.005 Landscape diversity (Shannon diversity of landscape elements) 3.5 1.63 0.033

85